1. Introduction
The share of renewables in the energy mix has seen a steep increase in the last decade, primarily driven by environmental concerns. In many regions of the world, the share of renewables is now at least 20% and is expected to increase during the coming decade to 40 or even 50% [
1,
2]. Recent developments in Li-ion battery modules that became much more competitive, as Li price on the world market continues to fall, have made it possible to efficiently store energy and release it to the grid in case of fluctuations in solar and wind power [
3]; this certainly has contributed to widespread acceptability of renewables by commercial stakeholders. Nevertheless, an efficient exploitation of renewables will also have to consider other energy storage possibilities, such as (super)capacitors and water electrolysis, to mention only the electrochemical methods. In particular, water electrolysis, mainly for the production of green hydrogen via the hydrogen evolution reaction (HER), is highly attractive, and could afford a powerful way for energy storage that is clean and sustainable and could be used even in remote and desertic regions.
HER is a well-known electrochemical interfacial process that essentially depends on pH, catalyst, and applied voltage. The reaction mechanisms and kinetics involved at the interface between electrolyte and catalyst have been described in detail in numerous review articles, e.g., article by Conway and Tilak [
4], Bockris and Conway [
5], and Eley et al. [
6], to mention only a few. The generally accepted reaction steps using a metal (M) catalyst under acidic conditions are [
4,
5,
6]:
followed by reactions (2) and/or (3)
Practically, the Tafel Equation (4) is used to evaluate catalyst’s performance.
where
is the hydrogen overpotential,
and
are temperature dependant materials constants, and
the measured current.
with
the gas constant,
the absolute temperature,
the Faraday constant, and
a constant.
is the exchange current density representing the equilibrium exchange current density of the forward and reverse reactions at the electrode. From Equation (4) we may infer that the larger
and
the smaller the overpotential
, a measure of a catalyst’s performance.
The mechanisms of HER have been rationalized by different authors in terms of the rate-determining mechanisms (Equations (1) to (3)), considering the different acting kinetics parameters such as adsorbed hydrogen surface coverage, overpotential and pH (e.g., Conway and Tilak [
4], Bhardwaj et al. [
7], Shingawa et al. [
8]). A critical parameter that has been largely discussed in the literature is the hydrogen adsorption energy,
, on the catalyst’s (C) surface which governs the hydrogen surface coverage and the rate-limiting step of diffusion. Together with the exchange current density,
, from the Tafel equation, E
C-H has been used to evaluate catalysts via the so-called volcano plot [
8,
9,
10,
11]. In simple terms, one may state that a moderate
and a higher
may constitute a meaningful materials selection criterium for catalysts development. To date, Pt has an unchallenged position at the top of the volcano, precisely because of the criteria mentioned above, and any new electrode material will be measured against the performance of Pt. However, Pt is a precious, strategic noble metal with limited resources, a fact that has triggered a spate of research for alternative materials. Several catalysts have been reported in the literature for the HER, including transition metals and various compounds, as compiled in the review article of Eftikhari [
9]. Catalyst performance is generally weighed in terms of the magnitude of overpotential necessary for HER at a given pH, the current density generated/weight of the active catalyst and durability/stability. As mentioned above, Pt that is mostly used in acidic conditions is well known for its outstanding performance [
12,
13,
14,
15]. In particular, Pt-NPs are the best known HER catalysts to date, performing at very low overpotentials. Noble metal-free catalysts such as Ni and Ni-Mo [
16], metal compounds, including FeS
2, MoS
2 [
17], as well metalloids compounds such as C
3N
4 [
18], to mention only a few of them, have been reported with acceptable performances, albeit stability issues have arisen. Also, while some of the catalysts mentioned above may perform well in tiny batches, their scalability is questionable as multiple processing steps are involved which complicates quality management and reproducibility of the results. New and efficient noble-metal-free HER catalysts that can be scaled-up using well established industrial processes are therefore of critical importance, as they will allow to cut electrode cost.
In the present work, the potential application of TiN as HER catalyst is considered. TiN is a well-established multifunctional material with applications spanning hard-coating for steel tools to plasmonics [
19,
20,
21]; it is usually processed in industrial scale as thin film of various thickness using physical (PVD) or chemical (CVD) vapor deposition (e.g., review articles [
22,
23]). Further, it has been shown using density functional theory that hydrogen adsorbs on TiN with the adsorption energy being dependent on termination (N or Ti) and adsorption sites. [
24,
25] The values obtained on TiN (111) range from 0.60 to 3.11 eV for N-termination and from 1.93 to 3.59 eV for Ti termination [
24]. A slightly different values were obtained by Siodmiak et al. on TiN (100) surface [
25]. Overall, the adsorption energies are considered to be moderate, albeit larger than on Pt. The diffusion barrier along the crystal sites was also shown to be rather low. This would unveil a promising non-noble metal electrocatalyst electrodes that are chemically and mechanically resistant. More advantages lie in the well-documented fact that they can be processed on an industrial scale using versatile and commercially available industrial equipment. It will be shown in this paper that TiN thin films of approximately 400 nm, deposited on Ti substrates via reactive gas sputtering, are potent HER catalysts in acidic media with current densities in the range of -70 mA/cm
2 at an overpotential of -0.1 V vs. SHE which is in the range of the performance of Pt/C. This performance is primarily ascribed to the existence of oxygen defects on the surface. It will also be indirectly shown that the TiN surface adsorption sites are critical for HER; few nanometres of sputtered Pd lead indeed to a drastic decrease in performance.
2. Materials and Methods
2.1. Synthesis and Characterization
Commercially pure, grade 1 Titanium sheets, 0.1 mm thick, were purchased in the annealed, oxide scale free and straightened condition from Goodfellow (Goodfellow, Germany). They were cut into 5x5 cm² samples, degreased in 99,9% ethanol in an ultrasonic bath, rinsed twice with ethanol and dried with compressed air, before introduction into the PVD chamber (PVD75, Lesker, Jefferson Hills, USA; electron beam system from Ferrotec GmbH, Unterensingen, Germany). An adhesion layer of 2 nm Ti was magnetron sputtered on the substrate surface using a Ti-target. On-top of the heterostructure above, the TiN layers were processed via reactive ion sputtering using 99.995 pure Titanium pellets (Lesker) and high-purity nitrogen (99.9999%, Westfalen AG, Germany). The following conditions were found to yield nearly stoichiometric TiN: electron beam evaporation of Ti at 8.75 kV and 65 to 90 mA in a mixture of nitrogen and Argon at a ratio of 8/15 and a pressure of 2 to 5x10-3 Torr. Under these conditions, the sputtering rate was in the range between 0.18 to 0.21 nm/s. The total sputtered thickness was 400 nm. The layer-thickness was cross-checked on cross-sections of coated Si that was introduced together with the Ti-substrates.
The microstructure and morphology of the samples were characterized with a high-resolution scanning electron microscope (SEM Ultra Plus, ZEISS, Oberkochen, Germany) operating in the secondary (SE) and energy selective backscattered (ESB) electron modes. The SEM is also equipped with an energy dispersive X-ray spectroscopy (EDS) package (INCAx-act, Oxford Instruments, High Wycombe, UK). The structure was characterized by X-ray diffraction (XRD, X’Pert Pro diffractometer PANalytical, Eindhoven, Netherland) in grazing incidence diffraction mode with constant θ = 1° using monochromatic CuKα radiation with λ = 1.5418 Å. The device has a full width to half maximum resolution of 0.03°.
An electrochemical workstation (ZAHNER IM6e, Kronach, Germany) was used for linear sweep voltammetry (LSV) measurements from -0.8 V to 0.3 V. The electrochemical experiments were performed in 0.5 M H2SO4 solution with pH 0.36, using a three-electrode set-up with a Pt mesh and HydroFlex (reversible H2 reference electrode) as counter and reference electrodes, respectively. The ratio of the working electrode (WE) area to that the counter electrode (CE) was approximately 1 to 4. All potentials were referenced to the reversible hydrogen electrode (RHE). The current was normalized by the sample’s measured area. All the H2SO4 solutions were saturated with Forming gas N2+H2 (after the usual bubbling with nitrogen) before measurement.
2.2. DFT Calculations
Using the Quantum ESPRESSO v.6.0 package [
26,
27,
28], ab initio calculations based on density functional theory were conducted to calculate charge distribution and hydrogen adsorption on 111- and 100-oriented TiN surfaces. In the present study, electro-ion interactions were described using projector augmented wave (PAW) potentials, while exchange correlations were represented using generalized gradient approximations (GGA) based on Perdew-Burke-Ernzerhof (PBE) frameworks [
29]. A 400 eV plane wave basis cut-off energy was set for the initial optimization of the structures by relaxing atomic positions using the Broyden-Fletcher-Goldfarb-Shanno (BFGS) method. To achieve high accuracy, a self-convergence field convergence criterion of 10-6 electrons was used. In
Figure 1, hydrogen adsorption is shown in various locations, namely top N, top Ti, bridge for TiN (100), and topTi, topN, bridge, hollow for TiN (111). Oxygen defects were introduced by replacing nitrogen atoms with oxygen, following the formula
, where x is set to 0.1. This specific concentration of oxygen ensures a calculated change in crystal structure that mainly affects the surface layer, while the bulk layers remain unchanged [
30]. This selective modification aims to explore changes in surface properties, such as reactivity and electronic structure, without significantly altering the overall properties of the material. The introduction of oxygen to the surface layer of TiN (111) represents a focused approach to modify material properties for targeted applications. To achieve the most stable surface hydrogen distance, the hydrogen atom and the top layer of the plate were allowed to relax their atomic positions. To calculate the hydrogen adsorption energies, the the usual definition of Equation (5) is used:
Figure 1.
Top view of modeled TiN(100) and TiN(111) with possible adsorption sites.
Figure 1.
Top view of modeled TiN(100) and TiN(111) with possible adsorption sites.
Figure 2.
top-view secondary electron (SE) micrographs of the TiN layers on Ti-substrate (a) and (b) which show at two different magnifications the particular morphology of the poly-nanocrystals with their faceted and rugged appearance. (c) is a cross-section micrograph obtained on a silicon substrate that was coated under the same conditions. Notice the perpendicular growth from high density nucleation sites at the interface (arrows).
Figure 2.
top-view secondary electron (SE) micrographs of the TiN layers on Ti-substrate (a) and (b) which show at two different magnifications the particular morphology of the poly-nanocrystals with their faceted and rugged appearance. (c) is a cross-section micrograph obtained on a silicon substrate that was coated under the same conditions. Notice the perpendicular growth from high density nucleation sites at the interface (arrows).
Figure 3.
SE micrographs of the Ti-TiN-Pd -layer at low (a) and high magnification (b) which show Pd-NPs topping the TiN crystal spikes. The right images show EDS analysis of Ti (c), TiN (d) and Ti-TiN-Pd (e).
Figure 3.
SE micrographs of the Ti-TiN-Pd -layer at low (a) and high magnification (b) which show Pd-NPs topping the TiN crystal spikes. The right images show EDS analysis of Ti (c), TiN (d) and Ti-TiN-Pd (e).
Figure 4.
XRD patterns of the substrate Ti, Ti/TiN and the Ti/TiN-Pd composite layer. The TiN pattern is indexed according to the PDF card # 01-087-0630 for the non-stoichiometric phase TiN0.88 (see below for discussion). The inset in the upper pattern is an enlargement of the pattern containing the 220 Pd reflex (*).
Figure 4.
XRD patterns of the substrate Ti, Ti/TiN and the Ti/TiN-Pd composite layer. The TiN pattern is indexed according to the PDF card # 01-087-0630 for the non-stoichiometric phase TiN0.88 (see below for discussion). The inset in the upper pattern is an enlargement of the pattern containing the 220 Pd reflex (*).
Figure 5.
XPS of Ti/TiN showing the core levels of Ti2p, N1s, O1s and C1s (see main text for more details). The measured Ti2p in (a) is deconvoluted in the green curve, showing bonding energies of the different compounds.
Figure 5.
XPS of Ti/TiN showing the core levels of Ti2p, N1s, O1s and C1s (see main text for more details). The measured Ti2p in (a) is deconvoluted in the green curve, showing bonding energies of the different compounds.
Figure 6.
(a) IR-corrected linear sweep voltammograms in 0.5 M H2SO4 of TiNO and TiNO-PdNPs. (b) Tafel plots corresponding to the LSV curves. The current density is normalised by the geometric surface area. Electrochemical impedance spectroscopy at different over potentials of -0.1, -0.2, -0.3 and -0.4 V for (c) TiNO and (d) TiNO-PdNPs. (e) Chronoamperometric curves in 0.5 M H2SO4 at -0.4 V for 16 hours.
Figure 6.
(a) IR-corrected linear sweep voltammograms in 0.5 M H2SO4 of TiNO and TiNO-PdNPs. (b) Tafel plots corresponding to the LSV curves. The current density is normalised by the geometric surface area. Electrochemical impedance spectroscopy at different over potentials of -0.1, -0.2, -0.3 and -0.4 V for (c) TiNO and (d) TiNO-PdNPs. (e) Chronoamperometric curves in 0.5 M H2SO4 at -0.4 V for 16 hours.
Figure 7.
Schematic illustration of the electronic changes, ions in the vicinity of the defect becoming more negative. Colour coded according to the change in atomic charge, with a colour gradient from positive (red) to negative (blue).
Figure 7.
Schematic illustration of the electronic changes, ions in the vicinity of the defect becoming more negative. Colour coded according to the change in atomic charge, with a colour gradient from positive (red) to negative (blue).
Figure 8.
Hydrogen adsorption energy profile of the HER on various sites of TiN(100), TiN(111) and (111).
Figure 8.
Hydrogen adsorption energy profile of the HER on various sites of TiN(100), TiN(111) and (111).
Figure 9.
Orbital electronic density of states for TiN(111) and TiN(100). EF denotes the Fermi level.
Figure 9.
Orbital electronic density of states for TiN(111) and TiN(100). EF denotes the Fermi level.
Figure 10.
Orbital electronic density of states for (111), and Hydrogen atoms adsorbed on different sites Ti, N, and O. H density of states are magnified by 100 for comparison. EF denotes the Fermi level.
Figure 10.
Orbital electronic density of states for (111), and Hydrogen atoms adsorbed on different sites Ti, N, and O. H density of states are magnified by 100 for comparison. EF denotes the Fermi level.
Table 1.
the values of Rs and Rct calculated using a reactive system as model.
Table 1.
the values of Rs and Rct calculated using a reactive system as model.
Overvoltage (V) |
Rs (Ohm) |
Rct (Ohm) |
Ti/TiN |
Ti/TiN/PdNPs |
Ti/TiN |
Ti/TiN/PdNPs |
-0.4 |
3.36 |
3.45 |
1.121 |
0.52 |
-0.3 |
3.48 |
3.32 |
1.845 |
0.76 |
-0.2 |
3.51 |
3.33 |
5.477 |
1.16 |
-0.1 |
3.52 |
3.35 |
38.97 |
2.18 |