Next Article in Journal
Superrepellent Porous Polymer Surfaces by Replication from Wrinkled Polydimethylsiloxane/Parylene F
Previous Article in Journal
Direct Observation of Evolution from Amorphous Phase to Strain Glass
Previous Article in Special Issue
Electroless Platinum Deposition Using Co3+/Co2+ Redox Couple as a Reducing Agent
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Three-Dimensional Au(NiMo)/Ti Catalysts for Efficient Hydrogen Evolution Reaction

by
Sukomol Barua
,
Aldona Balčiūnaitė
*,
Jūrate Vaičiūnienė
,
Loreta Tamašauskaitė-Tamašiūnaitė
and
Eugenijus Norkus
*
Department of Catalysis, Center for Physical Sciences and Technology (FTMC), LT-10257 Vilnius, Lithuania
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(22), 7901; https://doi.org/10.3390/ma15227901
Submission received: 31 August 2022 / Revised: 4 November 2022 / Accepted: 5 November 2022 / Published: 9 November 2022
(This article belongs to the Special Issue Advances in Electroless Metal Deposition)

Abstract

:
In this study, NiMo catalysts that have different metal loadings in the range of ca. 28–106 µg cm−2 were electrodeposited on the Ti substrate followed by their decoration with a very low amount of Au-crystallites in the range of ca. 1–5 µg cm−2 using the galvanic displacement method. The catalytic performance for hydrogen evolution reaction (HER) was evaluated on the NiMo/Ti and Au(NiMo)/Ti catalysts in an alkaline medium. It was found that among the investigated NiMo/Ti and Au(NiMo)/Ti catalysts, the Au(NiMo)/Ti-3 catalyst with the Au loading of 5.2 µg cm−2 gives the lowest overpotential of 252 mV for the HER to reach a current density of 10 mA·cm−2. The current densities for HER increase ca. 1.1–2.7 and ca. 1.1–2.2 times on the NiMo/Ti and Au(NiMo)/Ti catalysts, respectively, at −0.424 V, with an increase in temperature from 25 °C to 75 °C.

1. Introduction

Although fossil fuels, such as coal, oil, and natural gas, are the main energy sources and are widely used to meet energy needs, the increasing emissions of pollutants, carbon dioxide (CO2), and other greenhouse gases require the development of sustainable technologies to meet ever-increasing energy needs. Among various candidates to fulfill energy demands, hydrogen (H2) can be a potential substitute fuel for effective energy production and storage. H2 is a clean, economical renewable energy source and an excellent energy storage medium with excellent energy conversion efficiency, higher gravimetric energy density than gasoline (120 vs. 44 MJ kg−1), eco-friendliness, and zero carbon dioxide emission [1,2,3,4]. H2, an important chemical feedstock widely used in petroleum refining and ammonia synthesis, is industrially produced via coal gasification and steam reforming reaction under harsh conditions, resulting in the emission of greenhouse gases and micro-pollutants [5,6]. Among the various available methods, electrocatalytic water splitting is one of the most promising alternatives for H2 production that gained intense research interest in the last decades, as electricity-driven water splitting generates green H2.
It is well-known that H2 production by electrocatalytic water splitting in alkaline media is limited by the sluggish hydrogen evolution reaction (HER) kinetics and enormous electricity consumption. The HER mechanism of electrocatalytic water splitting includes three main steps, i.e., Volmer, Heyrovsky, and Tafel reactions, as shown below, in alkaline media [7]:
Volmer reaction: H2O + e → H* + OH
Heyrovsky reaction: H* + H2O+ e−→ H2 + OH
Tafel reaction: H* + H* → H2
Overall reaction: 2H2O + 2e → H2 + 2OH
The sluggish HER kinetics in alkaline media is mainly due to the fact that in the Volmer reaction, the proton source comes from the water molecule instead of H3O+ in the acid electrolyte, which involves additional energy to break the H–O–H bond [7]. To date, platinum (Pt) is the most effective and benchmark electrocatalyst for HER to achieve the lowest overpotential in both acidic and alkaline mediums, but unfortunately, because of the high production costs and the scarcity of resources, the use of Pt or other noble metals for electrodes for the water splitting process is not economically feasible [8,9]. In this context, it is the pursuit of most researchers to find an efficient, cost-effective and stable non-noble metal electrocatalyst for alkaline media to accelerate the Volmer step. Recent studies demonstrated that a number of non-noble transition metal-based materials, including nickel, molybdenum [10,11,12,13,14,15,16], cobalt [17,18,19], iron [20,21], tungsten [22,23], and transition metal compounds (TMCs), such as carbides [23,24,25], phosphide [26,27,28], nitrides [29], and sulfides [30], were investigated as electrocatalysts for HER to achieve excellent chemical stability due to their low cost and sufficient corrosion resistance under alkaline media. Additionally, low-cost transition metals, especially nickel-based electrocatalysts, received noticeable attention as supercapacitor electrodes and bifunctional electrocatalysts due to their abundant reserves, intrinsic high catalytic activity, excellent corrosion resistance, and high electrical conductivity [31]. A number of methods, such as spontaneous galvanic displacement [32], electrochemical deposition [13,33,34], hydrothermal synthesis, etc., were developed to explore and enhance the HER activity of Ni-based bi- and tri-metallic alloy catalysts Ni-M (M = Fe, Co, Mn, Mo, Cr, etc.) and Ni-M bimetallic oxides (BOs).
According to Engel–Brewer valence bond theory, whenever metals of the left half of the transition series (such as Ni and Co) are alloyed with metals of the right half of the transition series metals (Mo or, W), a synergistic effect can be anticipated in terms of hydrogen evolution activity [35]. The synergistic effect between Mo and Ni in the effect of hydrogen binding energy (HBE) is noteworthy, as the HBE between Ni and H is slightly weaker, whereas it is stronger enough between Mo and H. Therefore, the HBE can be controlled to a relatively moderate value by chemically coupling Ni and Mo, which can contribute to balancing the thermodynamics between hydrogen adsorption and desorption [36]. Moreover, enhanced HER activity demonstrated by a self-supported Ni–Mo–P ternary alloy coating on a three-dimensional (3D) Ni foam substrate (Ni–Mo–P/NF) were reported at a current density of −10 mA·cm−2 at a small overpotential of −63 mV in 1 M KOH electrolyte [13]. Heterostructured Ni–Mo–N composite nanoparticles, decorated on nitrogen-doped reduced graphene oxide (Ni–Mo–N/NG), also reported an excellent HER electrocatalytic activity with zero onset potential and 46.6 and 159.8 mV overpotentials for 10 and 100 mA·cm−2, respectively, in 1 M potassium hydroxide (KOH) solution [14].
This study presents a simple and low-cost procedure to fabricate efficient catalysts for HER. Three-dimensional (3D) binary Ni-Mo catalysts with different total metal loadings supported on a titanium (Ti) surface (denoted as NiMo/Ti) were prepared via the electrodeposition method, whereas for the decoration of the prepared NiMo/Ti catalysts with a small amount of Au crystallites, the galvanic displacement method was used.

2. Materials and Methods

2.1. Chemicals

Titanium foil (99.7% purity) and HAuCl4 (99.995%) were purchased from Sigma-Aldrich (Saint Louis, MO, USA) Supply. H2SO4 (96%), HCl (35–38%), nickel sulfate hexahydrate (NiSO4·6H2O, >98%), sodium molybdate dihydrate (Na2MoO4·2H2O, >99.5%), and NaOH (98.8%) were purchased from Chempur Company (Karlsruhe, Germany). Ultra-pure water with a resistivity of 18.2 MΩ·cm−1 was used for preparing the solutions. All chemicals were of analytical grade and used directly without further purification.

2.2. Fabrication of Catalysts

The catalysts were prepared by a facile, two-step process that involves electrodeposition of Ni2+ and Mo6+ ion on the surface of the Ti electrode, followed by a spontaneous Au displacement from the Au (III)-containing solution. Before the deposition of the NiMo catalysts, the Ti plates were pretreated in diluted H2SO4 (1:1 vol) at 70 °C for 3 s. NiMo catalysts were electroplated on the Ti surface (1 × 1 cm) from a bath containing 0.03 M Na2MoO4 along with 0.1, 0.2, and 1.0 M NiSO4 in an acidic condition (1.5 M H2SO4 and 1 M HCl). The chronopotentiometry was used for the NiMo coatings deposition on the Ti surface. The plating of coatings was carried out at the current of 0.1 mA and 1 mA for 3 min at each current. The Au crystallites were deposited on the prepared NiMo/Ti electrodes by their immersion into 1 mM HAuCl4 + 0.1 M HCl solution for 10 s. After plating, the samples were taken out, thoroughly rinsed with deionized water, and air-dried at room temperature.

2.3. Characterization of Catalysts

The morphology and composition of the prepared catalysts were investigated by scanning electron microscopy (SEM) TM 4000 Plus (HITACHI, Tokyo, Japan).
XRD patterns of pure Ti sheet, Ni-Mo/Ti, and Au-Ni-Mo/Ti catalysts were measured using an X-ray diffractometer D2 PHASER (Bruker, Karlsruhe, Germany). The measurements were conducted in the 2θ range of 10–90°.
The metal loadings were determined by inductively coupled plasma optical emission spectrometry (ICP–OES) analysis. The ICP–OES spectra were recorded using an Optima 7000DV spectrometer (Perkin Elmer, Waltham, MA, USA) at wavelengths of λNi 231.604 nm, λMo 202.031 nm, and λAu 267.595 nm.

2.4. Electrochemical Measurements

A conventional three-electrode electrochemical cell was used for electrochemical measurements. The fabricated NiMo/Ti and Au(NiMo)/Ti catalysts were employed as working electrodes, a Pt sheet was used as a counter electrode, and a calomel electrode was used as a reference. All potentials in this work were converted to the reversible hydrogen electrode (RHE) scale using the following Equation (5):
ERHE = ESCE + 0.242 V + 0.059 V × pHsolution.
Current densities were calculated using the electrodes’ geometric area of 2 cm2. Linear sweep voltammograms were recorded in a 1 M NaOH solution and always deaerated by argon (Ar) for 20 min prior to measurements. HER polarization curves were recorded from the open circuit potential (OCP) to −0.42 V (vs. RHE) at a polarization rate of 10 mV·s−1. Polarization curves were recorded at several temperatures from 25 to 75 °C, and temperatures were set with a water jacket connected to a LAUDA Alpha RA 8 thermostat. Stability was studied by recording chronoamperometry (CA) curves for HER at a potential of −0.22 V (vs. RHE) for half an hour. All electrochemical measurements were performed with a Metrohm Autolab potentiostat (PGSTAT302, Utrecht, The Netherlands) using the Electrochemical Software (Nova 2.1.4).

3. Results

In this study, we investigated the electrocatalytic activity of prepared 3D binary NiMo/Ti and ternary Au(NiMo)/Ti catalysts for HER. These catalysts were deposited on the Ti surface (1 × 1 cm) using an electrochemical bath. The optimal conditions for different 3D binary catalyst depositions were determined and are given in Table 1. The electrochemical deposition was carried out by applying the constant current of 0.1 mA and 1 mA for 3 min at each current. The Au crystallites were deposited on the prepared NiMo/Ti electrodes by their immersion into 1 mM HAuCl4 + 0.1 M HCl solution for 10 s at room temperature.
The morphology and composition of the prepared catalysts were studied by SEM. Figure 1 shows SEM images of the prepared different compositions 3D NiMo/Ti and Au(NiMo)/Ti catalysts. The low magnification SEM image of the 3D NiMo/Ti-3 catalyst (Figure 1c) proves the formation of many cedar leaf-like Ni-Mo alloy structures in a large area, in which leaf-like Ni-Mo alloy is uniformly dispersed on the Ti foil. These cedar leaf-like structures still retain much space among the leaves, forming a porous morphology that can be expected to facilitate electrolyte penetration. Many nanoparticles can be seen in Figure 1c, and they irregularly stack together, forming a cedar leaf-like structure. When Au crystallites were deposited on the prepared NiMo/Ti-3 electrode by being immersed into 1 mM HAuCl4 + 0.1 M HCl solution for 10 s, the porous leaf-like alloy structure was immediately covered with a tiny globular surface (Figure 1f). The mass of the elements (metal loadings) on the Ti substrate surface was determined by ICP–OES analysis (Table 2).
It can be seen that the formed 3D binary NiMo/Ti catalysts contained ca. 82.8–93.7 wt.% of Ni, whereas those 3D ternaries Au(NiMo)/Ti catalysts possessed 76.6–87.9 wt.% of Ni. The total metal loadings (µgmetalcm−2) in the prepared catalysts are quite different and vary from 23.9 up to 106.2 µgmetalcm−2. It should be noted that Ni and Mo amounts increase in the NiMo coatings by increasing the Ni2+ concentration in the plating solution, whereas the Mo amount was kept the same. Calculated Mo:Ni and Au:NiMo mass weight ratios are given in Table 2.
As seen from the data in Table 2, mass weight ratios Mo:Ni for NiMo/Ti increase due to the rise of Ni2+ concentration in the plating solution. A similar phenomenon is observed in the case of Au crystallite-modified NiMo/Ti catalysts. Mass ratios Mo:Ni also increase with the increase in the Ni2+ concentration in the plating solution (Table 2). Moreover, Au loadings in the AuNiMo/Ti-1, AuNiMo/Ti-2, and AuNiMo/Ti-3 catalysts increased while the deposition times of Au crystallites were the same—10 s. The mass ratio Au:NiMo slightly decreases. The increased Ni amount in the catalysts allows for achieving a higher Au loading in the ones.
Figure 2 shows XRD patterns for a pure Ti sheet (lower curve) and NiMo/Ti-3 and Au(NiMo)/Ti-3 catalysts (upper curves). Symbols indicate the positions of the XRD peaks of Ti (ICDD card no 00-044-1294).
The lowest XRD pattern (a) in Figure 2 contains sharp XRD peaks of the Ti sheet corresponding to the hexagonal structure of Ti. In the case of NiMo/Ti-3 and Au(NiMo)/Ti-3 catalysts, XRD peaks corresponding to Mo (110) and Mo (200) are shifted towards higher diffraction angles with respect to the positions of Mo presented in ICDD card no 00-044-1120. Additionally, the body-centered cubic lattice parameter decreases from 3.147 to 3.093 Å. This is the result of the formation of a solid solution of Ni (ICDD # 00-004-0850) with a small amount of Mo and Mo-Ni solid solution. There are no visible changes in the XRD patterns for NiMo/Ti-3 and Au(Ni-Mo)/Ti-3 (Figure 2, b,c curves) as the Au (ICDD # 00-004-0784) peaks can be amorphous or crystalline with low intensity.
The electrocatalytic properties of the prepared catalysts were investigated by recording LSVs in 1.0 M NaOH solution at a potential scan rate of 10 mV·s–1 in a potential range from open-circuit potential (OCP) up to −0.42 V (vs. RHE) for HER, at several temperatures from 25 up to 75 °C (Figure 3). Ternary Au(NiMo)/Ti-3 coating exhibited the highest current density (j), followed by Au(NiMo)/Ti-2 and Au(NiMo)/Ti-1, and the fabricated binary (NiMo/Ti) catalysts exhibited notably lower current density, in mutual comparison for HER (Figure 3). For those binary NiMo/Ti catalysts, the current density increases ca. 1.2–2.7 times with an increase in temperature from 25 up to 75 °C, whereas fabricated 3D ternary Au(NiMo)/Ti catalysts exhibit ca. 1.1–2.2 times higher current density for HER.
For instance, the current densities of −49.84, −40.73, and −36.58 mA·cm−2 were reached at −0.424 V (vs. RHE) using Au-decorated ternary Au(NiMo)/Ti-3, Au(NiMo)/Ti-2, and Au(NiMo)/Ti-1 catalysts, and relatively lower −34.81, −26.5, and −21.75 mA·cm−2 current densities were recorded at the same potential via using 3D binary NiMo/Ti-3, NiMo/Ti-2, and NiMo/Ti-1 catalysts at 25 °C, respectively (Figure 4a,b, Table 3). Overpotentials (vs. RHE) to reach the current density of 10 mA·cm−210) were found in a gradual increasing order for both Au(NiMo)/Ti and NiMo/Ti catalysts as follows:
Au(NiMo)/Ti-3 (−252 mV) < Au(NiMo)/Ti-2 (−298 mV) < Au(NiMo)/Ti-1 (−308 mV)
NiMo/Ti-3 (−288 mV) < NiMo/Ti-2 (−344 mV) < NiMo/Ti-1 (−349 mV).
It was determined that mass weight ratios Mo:Ni for NiMo/Ti catalysts increase due to the increased Ni2+ concentration in the plating solution (Table 2). A higher mass–weight ratio Mo:Ni induces a more pronounced activity of the NiMo/Ti-3 catalyst for HER. Additionally, the increased amount of Ni in the NiMo/Ti catalysts allows for achieving higher Au loading in the Au(NiMo)/Ti catalysts. This is the main factor that influences the lowering overpotential at Au(NiMo)/Ti catalysts compared with NiMo/Ti catalysts. The higher activity of Au crystallite-modified NiMo/Ti catalysts may be related with the synergetic effect of Au, Ni, and Mo [35].
HER polarization curves were then further used for constructing the Tafel plots and calculating the Tafel slope. Tafel slope values of 99.6, 100.5, and 130.4 mV·dec−1 were found for HER at NiMo/Ti-1, NiMo/Ti-2, and NiMo/Ti-3 catalysts, respectively. For those 3D ternary Au(NiMo)/Ti catalysts, Tafel slope values of 143.8, 98.7, and 131.2 mV·dec−1 were determined at Au(NiMo)/Ti-1, Au(NiMo)/Ti-2, and Au(NiMo)/Ti-3 catalysts, respectively (Figure 4a’,b’, Table 3). The determination of the Tafel slope explores the HER kinetics by measuring the increase in current density with the increase in overpotential, whereas the exchange current density (j0) reflects the electrode’s intrinsic activity for HER. The exchange current density (j0) was calculated for HER at all six catalysts by extrapolating the Tafel plots, η vs. log j. Thus, the j0 value of 0.144, 0.011, 0.076, 0.075, 0.006, and 0.005 mA·cm−2 were calculated for Au(NiMo)/Ti-3, Au(NiMo)/Ti-2, Au(NiMo)/Ti-1, NiMo/Ti-3, NiMo/Ti-2, and NiMo/Ti-1 catalysts, respectively (Table 3). It is worth noting that the j0 value determined for HER at the Au(NiMo)/Ti-3 coating was ca. 2–28 times higher than that determined for the rest of the studied catalysts.
Another crucial criterion for an advanced electrode material is its electrochemical stability. Chronoamperometric measurements with all six catalysts were carried out in 1 M NaOH at −0.22 V for 2 h. Initially, in the first 50–200 s, a decrease in current density was observed for all investigated catalysts. However, after approximately 500 s, the current densities settled down and remained apparently parallel throughout the experiments. CA results confirm the result of LSV analysis in terms of the ternary Au(NiMo)/Ti-3 catalyst, giving the highest current density during HER (−10.36 mA·cm−2 at 50 s) (Figure 5). A more than 2.5 times lower current density was obtained with Au(NiMo)/Ti-2 (−3.81 mA·cm−2) and ca. 5 times lower with Au(NiMo)/Ti-1 catalysts (−2.02 mA·cm−2). In the case of the binary NiMo/Ti-3 catalyst, a comparatively lower current density was recorded (−6.13 mA·cm−2 at 50 s) along with a 3–5 times lower value for NiMo/Ti-2 (−2.06 mA·cm−2) and NiMo/Ti-1 (−1.23 mA·cm−2) catalysts.
A comparison of HER parameters generated using herein-tested NiMo/Ti and Au(NiMo)/Ti catalysts in an alkaline medium with some electrodes reported in the literature is given in Table 4.

4. Conclusions

In summary, NiMo and Au(NiMo) catalysts supported on a titanium surface were studied as electrocatalysts for HER in an alkaline medium. NiMo/Ti catalysts with different total metal loadings in the range of ca. 28–106 µg cm−2 were prepared using a simple and low-cost metal electrodeposition method. The decoration of the prepared NiMo/Ti catalysts with a small amount of Au-crystallites in the range of ca. 1–5 µg cm−2 was carried out using the galvanic displacement method.
It was determined that, among the investigated catalysts, the Au(NiMo)/Ti-3 catalyst with the Au loading of 5.2 µg cm−2 exhibits the highest current density, as well as exchange current density during HER in a 1 M NaOH solution. Moreover, the Au(NiMo)/Ti-3 catalyst also displays excellent HER performance with an overpotential of 252 mV at a current density of 10 mA·cm−2.

Author Contributions

Conceptualization, A.B. and E.N.; methodology, J.V. and S.B.; formal analysis, S.B.; investigation, S.B. and J.V.; data curation, L.T.-T.; writing—original draft preparation, A.B. and S.B; writing—review and editing, E.N. and A.B.; visualization, L.T.-T. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by the European Social Fund under Measure No. 09.3.3-LMT-K-712-19-0138 ‘Development of Competences of Scientists, other Researchers and Students through Practical Research Activities’.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, W.; Xu, M.; Xu, X.; Zhou, W.; Shao, Z. Perovskite oxide-based electrodes for high-performance photoelectrochemical water splitting: A review. Angew. Chem. Int. Ed. 2020, 59, 136–152. [Google Scholar] [CrossRef] [PubMed]
  2. Ye, S.; Luo, F.; Zhang, Q.; Zhang, P.; Xu, T.; Wang, Q.; He, D.; Guo, L.; Zhang, Y.; He, C.; et al. Highly stable single Pt atomic sites anchored on aniline-stacked graphene for hydrogen evolution reaction. Energy Environ. Sci. 2019, 12, 1000–1007. [Google Scholar] [CrossRef]
  3. Wang, X.; Zhuang, L.; Jia, Y.; Liu, H.; Yan, X.; Zhang, L.; Yang, D.; Zhu, Z.; Yao, X. Plasma-triggered synergy of exfoliation, phase transformation, and surface engineering in cobalt diselenide for enhanced water oxidation. Angew. Chem. Int. Ed. 2018, 57, 16421–16425. [Google Scholar] [CrossRef]
  4. Wu, W.; Niu, C.; Wei, C.; Jia, Y.; Li, C.; Xu, Q. Activation of MoS2 basal planes for hydrogen evolution by zinc. Angew. Chem. Int. Ed. 2019, 58, 2029–2033. [Google Scholar] [CrossRef]
  5. Lin, L.; Sherrell, P.; Liu, Y.; Lei, W.; Zhang, S.; Zhang, H.; Wallace, G.G.; Chen, J. Engineered 2D transition metal dichalcogenides—A vision of viable hydrogen evolution reaction catalysis. Adv. Energy Mater. 2020, 10, 1903870. [Google Scholar] [CrossRef]
  6. Zhang, S.; Zhang, X.; Rui, Y.; Wang, R.; Li, X. Recent advances in non-precious metal electrocatalysts for pH-universal hydrogen evolution reaction. Green Energy Environ. 2021, 6, 458–478. [Google Scholar] [CrossRef]
  7. Luo, W.; Wang, Y.; Cheng, C. Ru-based electrocatalysts for hydrogen evolution reaction: Recent research advances and perspectives. Mater. Chem. Phys. 2020, 15, 100274. [Google Scholar]
  8. Chen, Y.; Zheng, Y.; Yue, X.; Huang, S. Hydrogen evolution reaction in full pH range on nickel doped tungsten carbide nanocubes as efficient and durable non-precious metal electrocatalysts. Int. J. Hydrogen Energy 2020, 45, 8695–8702. [Google Scholar] [CrossRef]
  9. Chao, T.; Hu, Y.; Hong, X.; Li, Y. Design of noble metal electrocatalysts on an atomic level. ChemElectroChem 2019, 6, 289–303. [Google Scholar] [CrossRef]
  10. Yan, Y.; Wang, P.; Lin, J.; Cao, J.; Qi, J. Modification strategies on transition metal-based electrocatalysts for efficient water splitting. J. Energy Chem. 2021, 58, 446–462. [Google Scholar] [CrossRef]
  11. Mohammed-Ibrahim, J.; Sun, X. Recent progress on earth abundant electrocatalysts for hydrogen evolution reaction (HER) in alkaline medium to achieve efficient water splitting—A review. J. Energy Chem. 2019, 34, 111–160. [Google Scholar] [CrossRef]
  12. Qian, G.; Chen, J.; Yu, T.; Liu, J.; Luo, L.; Yin, S. Three-phase heterojunction NiMo-based nano-needle for water splitting at industrial alkaline conditions. Nano-Micro Lett. 2022, 14, 20. [Google Scholar] [CrossRef] [PubMed]
  13. Toghraei, A.; Shahrabi, T.; Darband, G.B. Electrodeposition of self-supported Ni-Mo-P film on Ni foam as an affordable and high-performance electrocatalyst toward hydrogen evolution reaction. Electrochim. Acta 2020, 335, 135643. [Google Scholar] [CrossRef]
  14. Xue, S.; Zhang, W.; Zhang, Q.; Du, J.; Cheng, H.M.; Ren, W. Heterostructured Ni–Mo–N nanoparticles decorated on reduced graphene oxide as efficient and robust electrocatalyst for hydrogen evolution reaction. Carbon 2020, 165, 122–128. [Google Scholar] [CrossRef]
  15. Lu, X.; Cai, M.; Huang, J.; Xu, C. Ultrathin and porous Mo-doped Ni nanosheet arrays as high-efficient electrocatalysts for hydrogen evolution reaction. J. Colloid Interface Sci. 2020, 562, 307–312. [Google Scholar] [CrossRef]
  16. Mei, M.; Xu, X.; Wang, Y.; Wang, X.; Huo, Y. Three-dimensional supramolecular phosphomolybdate architecture-derived Mo-based electrocatalytic system for overall water splitting. Inorg. Chem. Front. 2018, 5, 819–826. [Google Scholar] [CrossRef]
  17. Liu, Y.; Yang, F.; Qin, W.; Yang, G. Co2P@ NiCo2O4 bi-functional electrocatalyst with low overpotential for water splitting in wide range pH electrolytes. J. Colloid Interface Sci. 2019, 534, 55–63. [Google Scholar] [CrossRef]
  18. Cai, Z.; Li, A.; Zhang, W.; Zhang, Y.; Cui, L.; Liu, J. Hierarchical Cu@ Co-decorated CuO@ Co3O4 nanostructure on Cu foam as efficient self-supported catalyst for hydrogen evolution reaction. J. Alloys Compd. 2021, 882, 160749. [Google Scholar] [CrossRef]
  19. Milikić, J.; Balčiūnaitė, A.; Sukackienė, Z.; Mladenović, D.; Santos, D.M.; Tamašauskaitė-Tamašiūnaitė, L.; Šljukić, B. Bimetallic Co-based (CoM, M= Mo, Fe, Mn) coatings for high-efficiency water splitting. Materials 2020, 14, 92. [Google Scholar] [CrossRef]
  20. Gong, Y.; Pan, H.; Xu, Z.; Yang, Z.; Lin, Y.; Wang, J. Crossed FeCo2S4 nanosheet arrays grown on 3D nickel foam as high-efficiency electrocatalyst for overall water splitting. Int. J. Hydrogen Energy 2018, 43, 17259–17264. [Google Scholar] [CrossRef]
  21. Liu, X.; Yin, Y.; Xiong, K.; Li, M. Facile synthesis of low-cost Fe3C-nitrogen and phosphorus co-doped porous carbon nanofibers: The efficient hydrogen evolution reaction catalysts. J. Alloys Compd. 2021, 856, 156213. [Google Scholar] [CrossRef]
  22. Lv, F.; Feng, J.; Wang, K.; Dou, Z.; Zhang, W.; Zhou, J.; Yang, C.; Luo, M.; Yang, Y.; Li, Y.; et al. Iridium–tungsten alloy nanodendrites as pH-universal water-splitting electrocatalysts. ACS Cent. Sci. 2018, 4, 1244–1252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Su, W.; Yan, P.; Wei, X.; Zhu, X.; Zhou, Q. Facile one-step synthesis of nitrogen-doped carbon sheets supported tungsten carbide nanoparticles electrocatalyst for hydrogen evolution reaction. Int. J. Hydrogen Energy 2020, 45, 33430–33439. [Google Scholar] [CrossRef]
  24. Huang, J.; Hong, W.; Liu, W. Molybdenum carbide nanosheets decorated with Ni(OH)2 nanoparticles towards efficient hydrogen evolution reaction in alkaline media. Appl. Surf. Sci. 2022, 579, 152152. [Google Scholar] [CrossRef]
  25. Kim, S.K.; Qiu, Y.; Zhang, Y.J.; Hurt, R.; Peterson, A. Nanocomposites of transition-metal carbides on reduced graphite oxide as catalysts for the hydrogen evolution reaction. Appl. Catal. B Environ. 2018, 235, 36–44. [Google Scholar] [CrossRef]
  26. Li, X.; Liu, Y.; Wu, Y.; Li, S.; Dai, Y.; Guan, J.; Zhang, M. Vanadium doped nickel cobalt phosphide as an efficient and stable electrode catalyst for hydrogen evolution reaction. J. Electroanal. Chem. 2021, 902, 115812. [Google Scholar] [CrossRef]
  27. Thi, M.L.N.; Tran, T.H.; Anh, P.H.; Nhac-Vu, H.T.; Bui, Q.B. Hierarchical zinc–nickel phosphides nanosheets on 3D nickel foam as self-supporting electrocatalysts for hydrogen evolution reaction. Polyhedron 2019, 168, 80–87. [Google Scholar] [CrossRef]
  28. Lu, Z.P.; Sepunaru, L. Electrodeposition of iron phosphide film for hydrogen evolution reaction. Electrochim. Acta 2020, 363, 137167. [Google Scholar] [CrossRef]
  29. Chen, P.; Ye, J.; Wang, H.; Ouyang, L.; Zhu, M. Recent progress of transition metal carbides/nitrides for electrocatalytic water splitting. J. Alloys Compd. 2021, 883, 160833. [Google Scholar] [CrossRef]
  30. Gao, T.; Nie, M.; Luo, J.; Huang, Z.; Sun, H.; Guo, P.; Xue, Z.; Liao, J.; Li, Q.; Teng, L. Nickel sulfides supported by carbon spheres as efficient catalysts for hydrogen evolution reaction. Electrochem. Commun. 2021, 129, 107076. [Google Scholar] [CrossRef]
  31. Wang, L.; Fan, J.; Liu, Y.; Chen, M.; Lin, Y.; Bi, H.; Liu, B.; Shi, N.; Xu, D.; Bao, J.; et al. Phase-modulation of iron/nickel phosphides nanocrystals “Armored” with porous P-doped carbon and anchored on P-doped graphene nanohybrids for enhanced overall water splitting. Adv. Funct. Mater. 2021, 31, 2010912. [Google Scholar] [CrossRef]
  32. Papaderakis, A.; Mintsouli, I.; Georgieva, J.; Sotiropoulos, S. Electrocatalysts prepared by galvanic replacement. Catalysts 2017, 7, 80. [Google Scholar] [CrossRef]
  33. Zhao, X.; He, B.; Zhang, J.; Du, C.; Ye, Q.; Liu, S. Electrodeposition of mesoporous Ni–Mo–O composite films for hydrogen evolution reaction. Vacuum 2022, 198, 110888. [Google Scholar] [CrossRef]
  34. Wu, Y.; He, H. Electrodeposited nickel–iron–carbon–molybdenum film as efficient bifunctional electrocatalyst for overall water splitting in alkaline solution. Int. J. Hydrogen Energy 2019, 44, 1336–1344. [Google Scholar] [CrossRef]
  35. Bodnarova, R.; Kozejova, M.; Latyshev, V.; Vorobiov, S.; Lisnichuk, M.; You, H.; Gregor, M.; Komanicky, V. Study of synergistic effects and compositional dependence of hydrogen evolution reaction on MoxNiy alloy thin films in alkaline media. Mol. Catal. 2022, 528, 112481. [Google Scholar] [CrossRef]
  36. Li, M.; Zhu, Y.; Wang, H.; Wang, C.; Pinna, N.; Lu, X. Ni strongly coupled with Mo2C encapsulated in nitrogen-doped carbon nanofibers as robust bifunctional catalyst for overall water splitting. Adv. Energy Mater. 2019, 9, 1803185. [Google Scholar] [CrossRef]
  37. Kuang, P.; Tong, T.; Fan, K.; Yu, J. In situ fabrication of Ni–Mo bimetal sulfide hybrid as an efficient electrocatalyst for hydrogen evolution over a wide pH range. ACS Catal. 2017, 7, 6179–6187. [Google Scholar] [CrossRef]
  38. Nady, H.; El-Rabiei, M.M.; Samy, M.; Deyab, M.A.; Abd El-Hafez, G.M. Novel Ni–Cr-based alloys as hydrogen fuel sources through alkaline water electrolytes. Int. J. Hydrogen Energy 2021, 46, 34749–34766. [Google Scholar] [CrossRef]
  39. Laszczyńska, A.; Tylus, W.; Szczygieł, I. Electrocatalytic properties for the hydrogen evolution of the electrodeposited Ni–Mo/WC composites. Int. J. Hydrogen Energy 2021, 46, 22813–22831. [Google Scholar] [CrossRef]
  40. Wang, J.; Wang, Y.; Xie, T.; Deng, Q. Facile and fast synthesis of Ni composite coating on Ti mesh by electrodeposition method for high-performance hydrogen production. Mater. Lett. 2019, 245, 138–141. [Google Scholar] [CrossRef]
  41. Wang, J.; Wang, Y.; Yao, Z.; Xie, T.; Deng, Q.; Jiang, Z.; Zhu, Q.; Liu, S.; Peng, Y.; Zhang, X. Enhanced hydrogen evolution activity of Ni/Ni3S2 nanosheet grown on Ti Mesh by Cu Doped Ni. J. Electrochem. Soc. 2019, 166, F168. [Google Scholar] [CrossRef]
  42. De Silva, U.; Masud, J.; Zhang, N.; Hong, Y.; Liyanage, W.P.; Zaeem, M.A.; Nath, M. Nickel telluride as a bifunctional electrocatalyst for efficient water splitting in alkaline medium. J. Mater. Chem. A 2018, 6, 7608–7622. [Google Scholar] [CrossRef]
  43. Ge, Y.; Gao, S.P.; Dong, P.; Baines, R.; Ajayan, P.M.; Ye, M.; Shen, J. Insight into the hydrogen evolution reaction of nickel dichalcogenide nanosheets: Activities related to non-metal ligands. Nanoscale 2017, 9, 5538–5544. [Google Scholar] [CrossRef] [PubMed]
  44. Yang, L.; Xu, H.; Liu, H.; Cheng, D.; Cao, D. Active site identification and evaluation criteria of in situ grown CoTe and NiTe nanoarrays for hydrogen evolution and oxygen evolution reactions. Small Methods 2019, 3, 1900113. [Google Scholar] [CrossRef]
  45. Nappini, S.; B’Olimpio, G.; Zhang, L.; Ghosh, B.; Kuo, C.-N.; Zhu, H.; Cheng, J.; Nardone, M.; Ottaviano, L.; Mondal, D.; et al. Transition-metal dichalcogenide NiTe2: An ambient-stable material for catalysis and nanoelectronics. Adv. Funct. Mater. 2020, 30, 2000915. [Google Scholar] [CrossRef]
Figure 1. SEM views of NiMo/Ti (ac) and Au(NiMo)/Ti (df) catalysts mentioned in Table 2. (c) The inset represents a photo of a cedar leaf.
Figure 1. SEM views of NiMo/Ti (ac) and Au(NiMo)/Ti (df) catalysts mentioned in Table 2. (c) The inset represents a photo of a cedar leaf.
Materials 15 07901 g001
Figure 2. XRD patterns of Ti sheet (a), Ni-Mo/Ti-3 (b), and Au(Ni-Mo)/Ti-3 (c) catalysts.
Figure 2. XRD patterns of Ti sheet (a), Ni-Mo/Ti-3 (b), and Au(Ni-Mo)/Ti-3 (c) catalysts.
Materials 15 07901 g002
Figure 3. HER polarization curves of 3D NiMo/Ti (ac) and Au(NiMo)/Ti (df) catalysts in 1 M NaOH solution at a 10 mV·s−1 potential scan rate and a temperature range (25–75 °C).
Figure 3. HER polarization curves of 3D NiMo/Ti (ac) and Au(NiMo)/Ti (df) catalysts in 1 M NaOH solution at a 10 mV·s−1 potential scan rate and a temperature range (25–75 °C).
Materials 15 07901 g003
Figure 4. HER polarization curves of 3D NiMo/Ti (a) and Au(NiMo)/Ti (b) catalysts in 1 M NaOH solution at a potential scan rate of 10 mV·s−1 and 25 °C temperature and corresponding Tafel plots (a’,b’).
Figure 4. HER polarization curves of 3D NiMo/Ti (a) and Au(NiMo)/Ti (b) catalysts in 1 M NaOH solution at a potential scan rate of 10 mV·s−1 and 25 °C temperature and corresponding Tafel plots (a’,b’).
Materials 15 07901 g004
Figure 5. Chronoamperometric data of the investigated NiMo/Ti and Au(NiMo)/Ti catalysts in 1 M NaOH solution at the potential value of −0.22V (vs. RHE), t = 2 h.
Figure 5. Chronoamperometric data of the investigated NiMo/Ti and Au(NiMo)/Ti catalysts in 1 M NaOH solution at the potential value of −0.22V (vs. RHE), t = 2 h.
Materials 15 07901 g005
Table 1. Composition of the electrochemical bath.
Table 1. Composition of the electrochemical bath.
CatalystsConcentration in Mol dm−3
Ni2+Mo6+
NiMo/Ti-10.10.03
NiMo/Ti-20.20.03
NiMo/Ti-31.00.03
Table 2. The metal loading in the catalysts was determined by ICP-OES analysis and metal mass weight ratio.
Table 2. The metal loading in the catalysts was determined by ICP-OES analysis and metal mass weight ratio.
CatalystNi Loadings
(µgNicm−2)
Mo Loadings
(µgMocm−2)
Auloadings
(µgAucm−2)
Total Metal Loading (µgmetalcm−2)Mass Weight Ratio
Mo:NiAu:NiMo
NiMo/Ti-123.44.9 28.31:4.78
NiMo/Ti-229.65.3 34.91:5.58
NiMo/Ti-399.56.7 106.21:14.85
Au(NiMo)/Ti-118.34.41.223.91:4.161:18.92
Au(NiMo)/Ti-225.44.61.731.71:5.521:17.65
Au(NiMo)/Ti-381.46.05.292.61:13.571:16.81
Table 3. Electrochemical performance of the tested catalysts toward HER in alkaline media.
Table 3. Electrochemical performance of the tested catalysts toward HER in alkaline media.
Catalystsj (mA·cm−2) in Different Temperatures (°C) at −0.424 VTafel Slope
(mV·dec−1)
η10 *
(mV)
j0
(mA·cm−2)
253545556575
NiMo/Ti-1−21.75−27.25−32.14−35.99−37.83−38.9199.6−3490.005
NiMo/Ti-2−26.5−31.94−39.72−49.53−55.44−61.05100.5−3440.006
NiMo/Ti-3−34.81−51.09−61.63−72.42−83.62−95.19130.4−2880.075
Au(NiMo)/Ti-1−36.58−39.32−47.04−54.93−61.86−68.68143.8−3080.076
Au(NiMo)/Ti-2−40.73−46.19−58.55−68.86−79.8−89.4598.7−2980.011
Au(NiMo)/Ti-3−49.84−56.38−68−78.81−91.6−102.86131.2−2520.144
* Overpotential at 10 mA cm−2.
Table 4. The electrochemical performance of herein tested catalysts towards HER in alkaline media and compared with that of transition metal-based electrodes reported in the literature.
Table 4. The electrochemical performance of herein tested catalysts towards HER in alkaline media and compared with that of transition metal-based electrodes reported in the literature.
CatalystOverpotential η10 ** (mV)Tafel Slope (mV·dec−1)Temperature
(°C)
ElectrolyteRef.
Au(NiMo)/Ti-3−252131.2251 M NaOHThis work
NiMo/Ti-3−288130.4251 M NaOHThis work
Au(NiMo)/Ti-2−29898.7251 M NaOHThis work
NiMo/Ti-2−344100.5251 M NaOHThis work
Au(NiMo)/Ti-1−308143.8251 M NaOHThis work
NiMo/Ti-1−34999.6251 M NaOHThis work
Ni-Mo-O MCFs−222.8141.6-1 M KOH[33]
NiFeCMo-30−254163.9-30% KOH[34]
NiS2/MoS2 HNW−20465-1 M KOH[37]
Ni–Cr–Mo–Fe,
Ni–Cr–Mo,
Ni–Cr alloy
−232
−249
−255
57.7
61.1
62.3
251 M KOH[38]
Ni–Mo/WC 1,
Ni–Mo/WC 2,
Ni–Mo/WC 3
−411
−262
−134
208
153
163
251 M KOH[39]
Ni/TM-360 s−205104-1 M KOH[40]
NiCu0.57/Ni3S2/TM,
Ni/Ni3S2/TM
−239
−441
86
195
-1 M KOH[41]
Ni3Te2-Ni foam
Ni3Te2-Au glass
Ni3Te2-Hydrothermal
−212
−237
−304
126.2
73.1
94.2
-1 M KOH[42]
NiTe2-nanosheet−25698-1 M KOH[43]
NiTeNR/NF−248185-1 M KOH[44]
NiTe2−520188.3-1 M KOH[45]
MCFs—mesoporous composite films, HNW—hybrid nanowire, WC—tungsten carbide, TM—Ti mesh, and NR/NF—nanorods/Ni foam. ** Overpotential at 10 mA·cm−2.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Barua, S.; Balčiūnaitė, A.; Vaičiūnienė, J.; Tamašauskaitė-Tamašiūnaitė, L.; Norkus, E. Three-Dimensional Au(NiMo)/Ti Catalysts for Efficient Hydrogen Evolution Reaction. Materials 2022, 15, 7901. https://doi.org/10.3390/ma15227901

AMA Style

Barua S, Balčiūnaitė A, Vaičiūnienė J, Tamašauskaitė-Tamašiūnaitė L, Norkus E. Three-Dimensional Au(NiMo)/Ti Catalysts for Efficient Hydrogen Evolution Reaction. Materials. 2022; 15(22):7901. https://doi.org/10.3390/ma15227901

Chicago/Turabian Style

Barua, Sukomol, Aldona Balčiūnaitė, Jūrate Vaičiūnienė, Loreta Tamašauskaitė-Tamašiūnaitė, and Eugenijus Norkus. 2022. "Three-Dimensional Au(NiMo)/Ti Catalysts for Efficient Hydrogen Evolution Reaction" Materials 15, no. 22: 7901. https://doi.org/10.3390/ma15227901

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