Hayk
Nersisyan
a,
Junmo
Jeong
b,
Hoyoung
Suh
c and
Jong Hyeon
Lee
*ab
aRASOM, Chungnam National University, 99 Daehak-ro, Yuseong-gu, Daejeon 34134, Republic of Korea. E-mail: jonglee@cnu.ac.kr
bGraduate School of Materials Science & Engineering, Chungnam National University, 99 Daehakro, Yuseong-gu, Daejeon 34134, Republic of Korea
cAdvanced Analysis Center, Korea Institute of Science and Technology, 5 Hwarang-ro 14-gil, Seonbuk-gu, Seoul 022792, Republic of Korea
First published on 15th November 2024
This study presents an efficient low-temperature process for synthesizing Mo nano- and microspheres for various applications. The synthesis process involves the preparation of a MoO3 + kZn mixture with an excess of zinc (Zn > 3) and processing to temperatures between 500 and 850 °C in an argon atmosphere. The growth kinetics of Mo particles are determined by analyzing the relationship between sphere diameter and processing time. Molybdenum nano- and microspheres are applied as electrocatalysts for the hydrogen evolution reaction (HER) and high electrocatalytic activity, including low overpotential (170–206 mV) and Tafel slope (40–50 mV dec−1) are recorded in 0.5 M H2SO4 electrolyte. DFT calculation provides adsorption Gibbs free energy for (001), (110), and (211) surfaces of Mo and charge density plots on pure Mo and Mo–O surfaces. As for vacuum-distilled Zn, its microstructure is also studied for its reuse and to assess its potential for additive manufacturing.
The electrocatalytic properties of Mo compounds, such as Mo2C and MoS2, make them suitable for applications for water splitting,4–8 as a host material in Na–S batteries,9 and N2 to NH3 conversion.10–12 The existing literature describes several methods for producing Mo micro- and nanopowders. These methods include the decomposition of Mo-hexacarbonyl complexes (Mo(CO)6),13 spray pyrolysis,14 mechanical milling,15 chemical vapor deposition,16 hydrogen reduction,17 radiofrequency plasma (RF) synthesis,18 carbothermal reduction19,20 molten salt synthesis,21 and combustion synthesis.22–27 While the existing methods have successfully produced Mo powder, the challenge is to prepare Mo microspheres at temperatures several times lower than the melting point of Mo (Tmelt. = 2610 °C). The desirable properties of Mo nano- and microspheres, such as good flow, packing, and controlled surface area, make it an ideal source for producing uniform bulk materials using conventional sintering processes and additive manufacturing techniques.
The plasma rotating electrode process (PREP) is a well-known synthesis technique that produces Mo microspheres ranging from 60 to 150 μm.28 The PREP method uses plasma, shielded by argon gas, to heat the end surface of a molybdenum rod above its melting point. Centrifugal forces separate the liquid Mo droplets from the rod and quickly cool them to solidify them into metal powder. The PREP method has some disadvantages, including the high cost of Mo powder and the need for specialized equipment. Ammonium paramolybdate ((NH4)6Mo7O24·4H2O) was used as a precursor to produce Mo microspheres using the RF hydrogen plasma method.29–31 During the synthesis, the (NH4)6Mo7O24·4H2O quickly evaporates. When the (NH4)6Mo7O24·4H2O vapor and hydrogen react in the gas phase, Mo microspheres with average sizes between 30 and 500 nm were obtained. Spray drying is an alternative method for producing Mo microspheres.32 The process involves drying an aqueous solution of (NH4)6Mo7O24·4H2O in hot nitrogen at 160 °C, then calcining at 500 °C and reducing with hydrogen at 950 °C for 2 hours. The Mo particles had diameters ranging from 5 to 15 μm, with a pitted surface observed in most of them. The only study in the existing literature that describes the synthesis of spherical Mo particles at temperatures significantly lower than Mo's melting point is the solvothermal/hydrothermal method.33 A mixture of (NH4)6Mo7O24·4H2O and ethylenediamine (C2H8N2) heated at 200 °C for 9 hours made nanospheres with an average diameter of 500 nm and specific surface areas of 2.31 m2 g−1. However, a significant amount of fine particles (less than 50 nm) are present on the surface of the nanospheres. The concentration of oxygen and the formation mechanism of nanospheres were not discussed.
This study demonstrated a novel and attractive approach for synthesizing Mo microspheres by heating a reactive mixture of MoO3 + kZn composition (where k is the number of moles of Zn) in argon atmosphere at temperatures between 500 and 850 °C. An excess of zinc (k > 3) used during the synthesis crates a liquid pool in which the nucleation and growth of Mo particles result in microspheres. Microspheres' growth kinetics and formation mechanism are described, and test experiments as electrocatalysts in 0.5 M H2SO4 electrolyte are conducted. Density functional theory (DFT) analysis was performed to analyze hydrogen adsorption energy on the (001), (110), and (211) surfaces of Mo. The morphology of vacuum-distilled Zn is also analyzed to assess its potential for reuse and additive manufacturing.
:
10 (excess Zn is 7 moles). The resulting mixture was then placed into an alumina boat and compacted by hand. The alumina boat containing the mixture was carefully sealed and positioned in a stainless steel pipe oven. A vacuum pump removed the air inside the pipe, and then the pipe was filled with argon gas at a pressure of 1 atm. Then, the pipe was heated to the selected temperature (from 500 to 850 °C) at a rate of 5 °C per minute, and the reaction mixture was held at maximum temperature from 0.17 to 20 hours. At the end of the heat treatment, a vacuum was also applied for 2 hours to facilitate the distillation of Zn. The distilled Zn was condensed on the colder sections of the reaction pipe and was collected for characterization. The main product was leached with a diluted solution of H2SO4:5H2O, washed with distilled water, and dried under a vacuum at a temperature range of 40–50 °C. As for the ZnSO4 solution, it can be used to recover Zn using a well-known commercial electrolysis technique.34,35 The process is carbon-free and environmentally clean.
To fabricate the working electrode (Mo/C/Nafion), 10 mg carbon + 90 mg Mo and 0.17 mL Nafion solution (5 wt%, DuPont, USA) were dispersed in 0.45 mL of normal propyl alcohol, followed by 440 G gravitational forced mixing using a centrifugal mixer (ARE-310, Thinky Mixer, Japan) for 20 min to form a homogeneous ink. Subsequently, five microliters of this ink was loaded onto a GCE (loading 1.4 mg cm−2) and dried overnight, after which liner scanning voltammetry (LSV) at a scan rate of 10 mV s−1 was used to record the polarization curves. The electrode potential was converted to the corresponding reversible hydrogen electrode potential (RHE) according to the equation ERHE = ESCE + 0.059 pH + 0.242 V, and the overpotential was calculated at a current density of 10 mA cm−2. The Tafel slope was determined using the Tafel equation: η = β
log(j + α), where β denotes the Tafel slope and α represents the Tafel parameter. Electrochemical impedance spectroscopy (EIS) analysis was conducted in the frequency range from 100 kHz to 0.001 Hz with a 10 mV amplitude. The electrochemical double-layer capacitances (Cdl) was estimated from cyclic voltammetry (CV) measurements in a non-faradaic region in 0.4–0.5 V range under different scan rates (25–150 mV s−1).
![]() | (1) |
E slab+H is the total energy of the slab with adsorbed hydrogen, Eslab is the energy of the slab, EH2 is the molecular hydrogen energy, and nH is the number of adsorbed hydrogen atoms. Here, to evaluate the oxide film's effect on the surface of Mo particles on hydrogen adsorption energy, Eslab refers to two cases: clean Mo surface and Mo surface with oxygen adsorbed. In hydrogen adsorption on catalytic surfaces, the Gibbs free energy (ΔG*) is a crucial parameter determining the feasibility and extent of adsorption under equilibrium conditions. The Gibbs free energy of adsorption is calculated as follows:
| ΔG* = ΔE* + ΔZPE − TΔS | (2) |
ΔZPE denotes the difference in zero-point energy, and TΔS reflects the change in entropy due to adsorption, which includes contributions from changes in the vibrational entropy of the system upon adsorption. This calculation assumes that the entropy change (ΔS) associated with hydrogen adsorption is negligible, so the calculated Gibbs free energy for the adsorption state can be corrected to ΔG* = ΔE* + 0.24 eV.37
| MoO3+ 3Zn = Mo +3ZnO + ΔHr | (3) |
| ΔHr = 3ΔHZnO − ΔHMoO3 = −71.72 kcal mol−1 | (4) |
Therefore, a DSC analysis was conducted to clarify the heat flow behavior in the MoO3 + 10Zn system with 7 moles of residual Zn. This residual amount of Zn was selected based on the results from preliminary experiments, particularly the reaction's heat and the shape of Mo particles.
As shown in Fig. 1(b), an exothermic event near the melting point of Zn (420 °C) occurred during the heating. This exothermic event rises due to the self-ignition of the mixture by the melting of Zn. A similar phenomenon for the WO3 + kZn system was recently observed and reported by our research team.38 The character of the second exothermic peak at 567 °C and subsequent endothermic peak at 582 °C is unclear. It is plausible to assume that the formation of the MoZn6 alloy phase and its melting may result in these events, as XRD analysis found this phase in the final products. The system's adiabatic temperature (Tad) and the equilibrium reaction phases (C) were determined by setting the initial reaction temperature as T* = 420 °C. The calculation was performed for the 3.0–12.0 range of k. After melting Zn, the system temperature rapidly increased due to the heat that evolved during the reduction reaction (Fig. 1(c)). At the stoichiometric point (k = 3), the Tad = 1243 °C; with increasing k (k = 10), the Tad decreases to 788 °C. The excess Zn (k > 3) consumes a particular part of the reaction heat for its melting and lowers the system temperature. However, even with excess Zn (k = 4–12), the system may react under the combustion regime and increase the system temperature to 700–1100 °C upon k. The main reaction phases predicted after the temperature spike are Mo, ZnO, Zn, and MoO2 (k < 4). For k less than 9 moles, the Zn may exist in both liquid and gaseous states; for k higher than 9 moles, the Zn is only in a molten state.
Experimentally, the thermal behavior of the MoO3 + 10Zn reaction near the melting point of Zn was investigated. The temperature of the reaction system was monitored by k-type thermocouples inserted into the pellet, as shown in the inset of Fig. 1(d). When the mixture temperature approaches the melting point of Zn, a shape increase to 800 °C occurs, as predicted by the thermodynamic analysis. The upper thermocouple (1) recorded this event first, and only after 14 seconds was it recorded by the bottom side thermocouple (2). This implies that the combustion reaction initiates at the top of the pellet and then self-propagates downwards, sequentially passing the first and second thermocouples. This phenomenon is known as a steady combustion process when the thermal wave self-propagates steadily from the pellet's top to bottom.39 The combustion wave velocity, calculated from the distance between the thermocouples (1.5 cm), is 0.11 cm s−1. The release of the low-speed combustion wave (instead of the volume combustion, when the reaction starts at each point of the sample simultaneously) dramatically decreases heat evaluation per unit of time and makes the process more controllable. The slight difference between T* and the Zn melting point could be attributed to the non-uniform heating of the sample or a standard error in temperature measurement.
The SEM micrograph in Fig. 2(b) shows the product after the combustion reaction. Two different morphologies can be distinguished: crystalline particles and molten fragments. The crystalline particles are ZnO as determined by EDS elemental analysis (Fig. 2(b, inset Table)), while the molten fragments contain Zn and MoZn6. After purification, the size of the purified Mo particle is determined to be less than 100 nm, as shown in Fig. 2(c). SEM/EDS spectral analysis revealed that the Mo content in the leached powder is only 80–85 wt%, and Zn and O impurities remained in it (Fig. 2(d)). Fig. S1 (ESI†) shows additional SEM images with different magnifications. Based on the analysis data, it can be postulated that even though the temperature rose to 800 °C after passing the combustion wave, the reaction time was insufficient to produce phase-pure and spherical Mo particles.
-The concentration of MoZn6 decreases significantly when the heat-treatment temperature exceeds 600 °C. XRD patterns of the products synthesized at 800 °C revealed only a tiny peak of MoZn6 at 2θ = 47°.
-The most intense peak, which includes the peaks of Mo and MoZn6 (2θ = 40.45°), also decreases as temperature increases. This is also caused by a reduction in MoZn6 concentration.
-A decrease in MoZn6 concentration led to increased Zn concentration (see peak intensity at 2θ = 43.3°).
When the mixture is thermally processed between 500 and 600 °C, it produces Mo and MoZn6 phases, similar to the product obtained after a combustion reaction. Above 600 °C, MoZn6 decomposes, resulting in Mo and Zn metals. Unfortunately, the literature does not provide the physical–chemical characteristics of MoZn6, including its melting, boiling, and decomposition temperatures.
Fig. 3(b) shows the XRD pattern of acid-purified samples produced at different temperatures. Despite the heat treatment temperature, the XRD peaks match the molybdenum. Meanwhile, the XRD peaks of Mo synthesized at 500 and 600 °C exhibit lower intensity than those of samples obtained at 800 °C. On the other hand, at 800 °C, the peak at 40.45° is very distinct and narrow and corresponds to the (110) plane of Mo.
Fig. 3(c–e) shows the scanning electron microscope (SEM) images of the purified products obtained at different temperatures.
When subjected to heat treatment at 500 °C and 600 °C, nanoparticles with a diameter of less than 100 nm were produced (Fig. 3(c)). At 700 °C, the SEM image displays nanospheres with a small portion of nanoparticles (Fig. 3(d)). The nanospheres have a diameter ranging from 0.2 to 0.5 μm, while the nanoparticles maintain a size of 100 nm or less. The nanospheres obtained at 800 °C exhibit increased diameters ranging from 1 to 3 μm (Fig. 3(e)). The microspheres are uniform and well-dispersed, although a small number of small particles can still be observed during the analysis.
The size of the Mo particles was not changed with further temperature increases up to 900 °C. This is due to the intense evaporation of residual Zn, which initiates at 850 °C, effectively preventing the Mo particles from growing any larger. This control over particle size is evident in the typical SEM micrographs of the Mo nano- and microspheres obtained at different processing temperatures (Fig. S2 (ESI†)).
The growth rate of Mo microspheres is directly influenced by the processing time. This was demonstrated through experiments conducted at 800 °C, with heat processing times ranging from 10 minutes (0.17 hours) to 1200 minutes (20 hours). The SEM images of purified samples (Fig. 3(f–h) and Fig. S3 (ESI†)) clearly show that the size of the Mo particles consistently increases with processing time, reaching 3–8 μm after 20 hours. As the heating time increases, the surface of the particles becomes very clean without any small particles. Merging two or more smaller microspheres to form a larger one is a notable observation, as shown in Fig. 3(g).
Fig. 4A(a and b) shows the micrographs of Mo spherical particles obtained from the MoO3 + 10Zn mixture at 800 °C for 2 hours. The particles have a perfectly spherical shape and are well dispersed. SEM/EDS mapping shows the presence of two elements, Mo and O. The oxygen is primarily found on the surface layer of the Mo particles, and its concentration varies from 0.5 to 2.5 wt% depending on the size of the microspheres (Fig. 4A(c)). When the thermal processing time is increased beyond 10 hours, the oxygen content reduces to below 0.3 wt% (Fig. S4, ESI†).
TEM analysis of Mo synthesized at 800 °C for 2 hours revealed spherical Mo particles with small particles attached to their surface (Fig. 4B(a and b)).
These tiny particles contain Mo, Zn, and O. The SAED pattern in Fig. 4B(c) shows (1
0), (1
2), and (002) planes of Mo in the [110] direction. Moreover, TEM/EDS mapping of a microsphere demonstrates that the oxygen distribution on its surface is not uniform, potentially due to small particles partially attached to the microsphere's surface (Fig. 4B(d)).
![]() | ||
| Fig. 5 (a) Raman spectra of Mo powder (800 °C); (b) adsorption/desorption isotherms of Mo (500 °C); (c) full XPS spectra of Mo powder (800 °C); (d) Mo spectra with Gaussian fitting. | ||
The adsorption–desorption isotherms for Mo derived at 500 °C are shown in Fig. 5b. This isotherm resembles type-II, which is highly common in physical adsorption in powdered samples. The typical hysteresis loop (H3 type) between the desorption and adsorption branches indicates the sample has macroporous structure, which may arise due to the agglomeration of particles. The inset of Fig. 5b shows the BJH pore volume according to pore diameter. According to the inset, the Mo exhibit a broad size distribution, reaching a maximum at 47 nm, thus indicating the mesoporous-macroporous nature of Mo nanoparticles. The following BET surface area values were obtained for Mo samples synthesized at various temperatures: 20.58 m2 g−1 (Mo-500 °C), 5.34 m2 g−1 (Mo-700 °C), 1.75 m2 g−1 (Mo-800 °C, 2 hours) and 0. 73 m2 g−1 (800 °C 20 hours). The oxygen concentration in Mo powder determined by Eltra ONH-2000 O/N analyzer shows strong decreasing tendency with processing time: 1.82 wt% (800 °C, 2 h), 0.74 wt% (800 °C, 10 h) and 0.35 wt% (800 °C, 20 h).
XPS analysis was conducted on the Mo microspheres obtained from the MoO3 + 10Zn mixture after 2 hours of treatment at 800 °C (Fig. 5(c)). The most peaks identified on the Mo particle surface are Mo 3d, O 1s, Zn 2P, and C 1s. Mo 3d peak after Gaussian fitting shows three peaks with 228.12337, 232.08, and 235.10075 eV binding energies42 (Fig. 5(d)). These peaks correspond to Mo6+ (3d5/2), Mo6+ (3d3/2), and Mo4+ (3d5/2) electronic states, confirming the presence of MoO3 and MoO2 on the particle surface. Moreover, besides molybdenum oxides, Zn 2p3/2 and Zn 2p1/2 states possibly indicate the presence of Zn-containing phases (Fig. S5 (ESI†)). Based on the XPS analysis data, we can infer that the presence of Zn 2p3/2 and Zn 2p1/2 states is attributed to small particles with ZnMoxOy composition. XPS analysis also detected a C 1S peak at 284.0 eV and an N 1s peak at 398.08 eV (Fig. S5 (ESI†)). The C 1s peak is associated with sp2 hybridized carbon, but its origin is unclear. Similarly, the origin of the N 1s peak is also ambiguous (Fig. S5 (ESI†)). The argon gas used in the experiments may source these impurities.
![]() | ||
| Fig. 6 Growth tendency of Mo spheres: (a) upon the processing temperature; (b) upon the heating time; (c) data presented in D3 − Do3/t coordinates; (d) Mo diameter-growth rate relation. | ||
The impact of the holding time at 800 °C on the sphere's diameter was also examined and depicted in Fig. 6(b). This distribution can be divided into two zones: 1st zone (0–1 hours), where the particle growth has a linear character, and 2nd zone (2–20 hours), where the particle diameter growth takes a parabolic character. To calculate particle growth kinetics based on the diameter–time relationship, we utilized the standard model from the Lifshitz–Slyozov–Wagner (LSW) theory, which describes the growth kinetics of particles during Ostwald ripening.43 According to the LSW theory, the relationship between particle diameter and time can be characterized by:
| D3 − Do3 = Kt | (5) |
| γ = K/3D2 | (6) |
By eqn (6), we calculated the growth rate of microspheres based on their size (Fig. 6(d)). It is evident that as the diameter of a microsphere increases from 0.5 to 10 μm, the growth rate decreases by almost two orders of magnitude.
The Mo particle has a unique spherical shape maintained during coarsening through the Ostwald ripening process. The liquid Zn medium drives the formation of spherical particles. The nature of the fluid medium and van der Waals interaction between the liquid and nanoparticles play a critical role in particle aggregation, attachment, assembly, and shaping.44,45 In the Mo–Zn system, two main factors are essential: the melting of Zn and forming MoZn6 alloy. In this scenario, the growth of the Mo microsphere can be described as a complex process with several steps. The interfacial morphologies are primarily influenced by diffusion in the liquid and solid phases.
A thermochemical scheme for the growth process of the Mo microsphere can be suggested as follows:
-Reduction of MoO3 and Mo nuclear formation (500–600 °C)
| MoO3(sol.) + 10Zn(liq.) → Mo (sol.) +3ZnO (sol.) + 7Zn (liq.) | (7) |
-MoZn6 alloy phase formation (600–650 °C)
| Mo (sol.) + 7Zn (liq.) → MoZn6 (liq.) + Zn (liq.) | (8) |
-Decomposition of MoZn6 (650–750 °C)
| MoZn6 (liq.) → Mo (sol.) + 6Zn (liq.) | (9) |
-Growth of microsphere by liquid phase sintering (750–850 °C)
| Mo(sol.) + Zn (liq.) → Mo (sol.) (larger diameter) + Zn(liq.) | (10) |
During the liquid phase sintering, liquid zinc helps reduce particles' surface energy, allowing them to come together and form larger ones. The examination under the SEM shows that the growth of micrometer-sized microspheres mainly happens when two or more similar particles merge to form a larger one. This merging process, facilitated by the presence of liquid zinc, resembles the mechanism of liquid phase sintering. As evidence of this mechanism, SEM/EDS analysis always reveals the presence of 1–3 wt% zinc in molybdenum (Mo) microspheres before vacuum distillation. However, following vacuum distillation, the zinc completely evaporates from the microspheres, while the spherical shape of the particles remains intact.
It is important to note that a similar study involving the WO3 + kZn system produced round-shaped tungsten (W) nanoparticles measuring 100 nm in size (refer to Fig. S6, ESI†). However, unlike the Mo-system, there was no increase in particle size and further spherodization. A similar result was obtained from the MoO3 + 30 Mg mixture at 800 °C. As demonstrated in Fig. S7 (ESI†), the growth of Mo particles in liquid Mg did not result in the formation of nano- and microspheres, yielding only Mo nanoparticles.
This suggests the significance of intermediate MoZn6 alloy phase formation and its important role in Mo particles spherodization process.
To gain more insight into the electrode kinetics of Mo catalysts under the HER conditions, EIS measurements on the catalysts were conducted over the same potential range and pH 0.5. As shown in Fig. 7(c), Nyquist plot of each catalyst was measured using their overpotentials at current density 10 mA cm−2. The Rct values decrease from Mo-800 to Mo-500, conforming that the catalytic activity of Mo samples increases with increasing the synthesis temperature.
CV tests in non-faradaic regions was also conducted at various scan rates (25–150 mV s−1) to estimate the double layer capacitance (Cdl) of Mo catalysts. Fig. 7(d) illustrates the measured representative CVs for the Mo-800 catalysts (see Fig. S9 in ESI† for corresponding CVs for Mo-700 and Mo-500). The Cdl was determined from CV using the equation:47
| Cdl = Δj(ja − jc)/2ν | (11) |
Faradaic efficiency of Mo-catalysts was calculated following the protocols described in49 and described in ESI† (Section 7, Fig. S10 and S11). The H2 generating rate is calculated as 0.61 mL min−1, and the faradaic efficiency approaches nearly 83.7% at 2.82 V. For comparison the faradaic efficiency of commercial Pt/C electrocatalyst is 81.98–85.97%, at 1.5–1.8 V respectively.50 This means that Mo/C electrocatalysts demonstrate sufficiently high faradaic efficiency.
Mo and Mo-supported MoS3 electrocatalysts on a FTO substrate for HER in 0.5 M H2SO4 solution demonstrated the following overpotentials: FTO/Mo – 326 mV, FTO/α-MoS3 – 520 mV, and FTO/Mo/α-MoS3 – 247 mV.51 Other non-noble metals as electrocatalysts for HER are rare, and there is limited data in the literature.52 Among the non-noble metals, Ni is the most well-known electrocatalyst, so a comparison can be made between Ni and Ni–Mo electrocatalysts. Ni has a lower hydrogen adsorption free energy (ΔGH+ = −0.15 eV) than Mo (ΔGH+ = −0.38 eV),53 and consequently, it is expected to show better electrochemical performance. However, pristine Ni has a very high overpotential (>500 mV), and the overpotential of pristine Mo–Ni electrode is about 270–280 mV.52 The significantly low overpotential of Mo-800 (170 mV) indicates that the current synthesis process may produce Mo spheres with high surface activity and improved catalytic performance for water splitting.
Note that the Tafel slopes of Mo samples at 10 mA cm−2 current density are almost comparable with the commercial Pt/C catalyst.54 We believe further research on this material can be attractive to lower the overpotential and fabricate an inexpensive and effective catalyst for water splitting.
The calculation results for the planes are depicted in Table 1.
| Mo slab surface conditions for H adsorption | ΔGH*, adsorption Gibbs free energy, eV | |
|---|---|---|
| (001) | On clean Mo | −1.0254 |
| On oxygen-adsorbed Mo | −0.7554 | |
| (110) | On clean Mo | −0.9935 |
| On oxygen-adsorbed Mo | −1.1254 | |
| (211) | On clean Mo | −1.683 |
| On oxygen adsorbed Mo | −1.0761 | |
The calculated Gibbs free energy of hydrogen adsorption for clean Mo slab immensely differs from the |0.2| eV typically reported for suitable catalysts. This indicates that metal Mo has a strong interaction with hydrogen and is easy to adsorb. Still, it may be difficult to escape into the gas phase after hydrogen molecules are formed.
Among the three planes considered in the calculation, the hydrogen adsorption energy of the (211) plane has the most significant negative value, so the contribution of Mo to the catalytic properties is not expected to be significant. Notably, the hydrogen adsorption energy of Mo in the presence of a surface oxide layer decreases drastically for facets (001) and (211).
Fig. 8(b) illustrates the differential charge density plots of stable hydrogen (H) adsorption configurations on a molybdenum surface with and without an oxide layer. The yellow area surrounding the atom denotes an augmentation in the charge density of the bonding region. In contrast, the blue area signifies a reduction in the charge density of the bonding region.
As shown in Fig. 8(b), a charge transfer occurs between Mo atoms and hydrogen atoms on a pure Mo surface, decreasing charge density at the lower part of the surface of Mo atoms. Conversely, in the presence of an oxide, the Mo–O bond is stable, and it can be observed that electron transfer occurs between oxygen atoms and hydrogen atoms. In other words, when hydrogen is stably adsorbed on the surface of oxidized Mo, Mo, and oxygen atoms in the adsorption system, it loses some electrons, and the charge density in the Mo–O bonding region decreases, thus maintaining the Mo–O bond. It is hypothesized that the interaction between Mo and O facilitates higher adsorption energy for hydrogen than that observed on a pristine Mo surface, except the (110) plane. This is postulated to be due to the aligned and dense arrangement of Mo atoms on the (110) plane, which does not exhibit an offsetting effect on the adsorption energy by the surface oxide layer.
0}, and {11
1}, can be seen.
In summary, while spherical particles have lower surface energy, the formation of a crystal structure is favored during crystallization because it leads to a lower total Gibbs free energy of the system. The internal energy savings from the ordered crystal lattice outweigh the costs associated with the increased surface energy of the faceted crystal. The transition to faceted crystals is likely to occur during the cooling near the Zn melting point, similar to the behavior of gold particles, as reported in.55–57
Information about DFT calculation, including calculation codes and video files, will be presented by request.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4qm00814f |
| This journal is © the Partner Organisations 2025 |