Open Access Article
0) surface
Zezhong Miaoab,
Xing Zhuc,
Yuqian Jind,
Lingzhao Kong*c and
Shenggang Li
*ab
aLow-Carbon Conversion Science and Engineering Centre, and, State Key Laboratory of Low Carbon Catalysis and Carbon Dioxide Utilization, Shanghai Advanced Research Institute, Chinese Academy of Sciences, Shanghai 201210, China. E-mail: lisg@sari.ac.cn
bSchool of Physical Science and Technology, ShanghaiTech University, Shanghai 201210, China
cSchool of Environmental Science and Engineering, Suzhou University of Science and Technology, Suzhou, Jiangsu 215009, P.R. China. E-mail: konglz@sari.ac.cn
dDepartment of Chemistry, University of California, Irvine, Irvine, CA 92697, USA
First published on 23rd December 2025
ZnO, an important component in many catalysts for the hydrogenation of carbon monoxide and carbon dioxide, upcycling of plastics and hydrodeoxygenation of biomass, exhibits a strong capacity for H2 activation. This work examines eleven distinct H2 activation pathways on pristine and defective ZnO (10
0) surfaces, demonstrating that the OV–Zn3 ensemble is not a spectator site. Instead, OV–Zn3 acts as an electron reservoir with strong electron-donating ability, albeit with limited electron-storage capacity. This region interacts with surface H adsorbates and, while modulating the behavior of the adsorbed H species, undergoes lattice distortion and electronic rearrangement as the adsorption sites vary. Furthermore, the tendency of the H atoms to adsorb on the Zn–O pairs drives the growth of a one-dimensional H-chain along the [0001] direction, leading to distinct diffusion behavior along the [0001] and [1
10] directions. The existence of multiple H2 activation routes and H diffusion pathways provides a rational explanation for the experimentally observed variations in the OV concentration as well as the hydrogen coverage at the OV sites. By correlating these atomic-scale insights with available experimental observations, we propose how defect engineering and thermal control could be synergistically employed to tune H2 activation on ZnO surfaces, providing a fresh perspective for rational catalyst design of ZnO-based hydrogenation catalysts.
12 are widely used for CO2 hydrogenation to methanol, syngas conversion into light olefins and the water–gas shift reaction. ZnO-based catalysts also play a key role in plastic upcycling13–15 and biomass conversion via hydrodeoxygenation.13,16,17 Despite major advances, there remain uncertainty regarding the reaction mechanism owing to the structural complexity of the catalysts and the multitude of adsorbates and intermediates involved.
Recent studies indicate that the ZnO component in these catalysts often plays a decisive role in H2 activation. For Cu/Zn/Al2O3, which is the mature industrial methanol synthesis catalyst,18 the Cu/ZnO interface or CuZn alloy has been widely recognized as the primary active site.2,19 H2 molecules are believed to dissociate at the ZnO/Cu interface with a relatively low energy barrier and subsequently spillover to terminal oxygen atoms on the ZnO surface, contributing to oxygen vacancy (OV) formation.20 For the ZnZrOx solid solution catalyst, Zn–O pairs have been identified as the main active sites for heterolytic H2 dissociation.21,22 Similarly, on the surface of ZnAl2O4 spinel catalysts, the formation of an amorphous ZnO phase is believed to facilitate H2 activation, thereby enhancing the methanol production rate.23 Effective H2 activation not only provides the necessary hydrogen for subsequent reactions but also promotes the formation of OVs, which can serve as highly active sites for CO2 activation.24
Liu et al. constructed an inverse Ni–ZnO interface with interfacial frustrated Lewis pairs capable of heterolytically cleaving H2 into Hδ+/Hδ− and thereby accelerating carbonyl hydrogenation.25 A Ni1Fe1–ZnO interface was engineered to enable quantitative conversion of methyl stearate by hydrogenation to octadecanol with 92.7% selectivity and an initial rate three times higher than that of NiFe/C.26 In addition, for polyol hydrogenolysis (e.g., glycerol to 1,2-propanediol), ZnO in the classic Cu–ZnO catalytic system serves as an “atomic-hydrogen reservoir”, which promotes hydrogen spillover, as it can promote both H2 activation and the hydrogen transfer steps required for selective hydrogenolysis.27 ZnO by itself possesses an inherently high capacity for H2 activation, which makes it well suited for use as an H2 gas sensor28–30 for detecting leaks in hydrogen fuel storage systems.31 Even at temperatures as low as 20 K, H2 can dissociate and form one-dimensional (1D) hydrogen chains aligned along the [0001] direction on the ZnO (10
0) surface.32 Infrared study33 and theoretical calculations34,35 have demonstrated that H2 activation on the ZnO (10
0) surface proceeds via heterolytic dissociation at Zn–O pairs, forming a hydride (H–Zn) and a hydroxyl (O–H).
Activated hydrogen species can readily diffuse on the ZnO (10
0) surface, both along [0001] and [1
10]. Hydrogen atoms adsorbed on Zn have been observed to migrate to adjacent O atoms.34,36 However, DFT calculations34 indicate that migration to neighbouring O4c and O3c sites has reaction energies of 0.09 eV and −0.01 eV, respectively, with activation barriers of 1.40 eV and 1.10 eV. The final states are not thermodynamically more stable than the initial states, and the kinetic barriers are relatively high, which together do not strongly support this scenario. Recently, Ling et al. observed hydrogen diffusion along the [1
10] direction,37 where the atomic rows are composed solely of either Zn or O atoms (Fig. 1a) rather than alternating Zn–O pairs. However, the driving force and mechanism governing the diffusion across identical atomic sites remain unresolved.
OVs have also been experimentally identified on ZnO surfaces.38 Although Ling et al. previously proposed that H2 dissociation on ZnO surfaces proceeds without the assistance of OVs,37 recent studies indicate that the contribution of these defects to H2 activation remains largely overlooked. Song et al. exposed ZnO to H2 at 200 °C and detected a single resonance at 8.4 ppm in the 1H MAS NMR spectrum.34 They unequivocally attributed it to hydrogen species adsorbed at OVs. This finding provides evidence for the probability of H2 dissociation at these defect sites. Subsequent experiments36 tracked how the coverages of OVs and hydrogen species on the ZnO (10
0) surface evolve with temperature under an H2 atmosphere. At 40 °C, heterolytic H2 dissociation on the stoichiometric surface dominated, with no OVs observed, and at 150 °C, vacancies began to form, but no hydrogen signal associated with them was detected. At 200 °C, more OVs were generated from a deeper reduction, which became populated by hydrogen, so the net vacancy concentration did not rise markedly relative to that at 150 °C, while at 300 °C, both the OV density and the coverage of vacancy-bound hydrogen decreased, ascribed to migration of lattice oxygens from the bulk to the surface.
In this work, we elucidate the mechanism by which an OV–Zn3 motif forms on ZnO (10
0) surfaces. We further examine eight distinct chemisorption configurations for atomic hydrogen, and eleven for molecular H2 on the ZnO surface, to probe multiple modes of interaction between hydrogen species and the surface, as well as their impact on lattice distortion, which is a powerful lever for tuning catalytic performance.39 In addition, we investigate the self-assembly of H species into 1D hydrogen chains and their surface diffusion. Viewed from the behaviour of H species, our analysis offers a possible explanation for the experimentally observed temperature dependence of OV and hydrogen coverages. These insights are important for the rational design of more effective ZnO-based catalysts.
Crystal orbital Hamilton population (COHP) analyses were performed with the LOBSTER package.51 For each atom pair A–B, the integrated COHP (ICOHP) was computed as:
Lattice parameters of the primitive unit cell (PUC) of bulk ZnO (Fig. S1a) were calculated as a = b = 3.29 Å and c = 5.30 Å, compared favourably with their experimental values of a = b = 3.25, and c = 5.21 Å.53 For the ZnO (10
0) surface, a periodic supercell of (3 × 2) with 3 repeated layers along the z direction was constructed (Fig. S1b and e). To eliminate spurious interactions between periodic images, a vacuum region of 15 Å was introduced along the z-direction. A Γ-centered Monkhorst–Pack54 k-point mesh of (3 × 3 × 1) was employed to sample the Brillouin zone, and a denser (5 × 5 × 1) grid was used for electronic-structure calculations.
The formation energy of a thermally induced OV is defined as the reaction energy for the thermal desorption of molecular O2:
In addition, the formation energy of a hydrogen-induced OV, with respect to gas-phase H2/H2O, is defined as:
| ΔEf,h-OV = EOVslab − Epristineslab − EH2 + EH2O |
In the above equations, EOVslab and Epristineslab are the total energies of the defect and pristine slabs. EO2, EH2, EH2O are the energies of gas-phase O2, H2 and H2O, respectively.
The adsorption energy of adsorbate A on a slab is defined as:
| Ead,A = Etotal − (Eslab + EA) |
Here, Etotal is the total energy of the adsorbate–slab system, Eslab is the energy of the clean slab, and EA is the energy of the isolated adsorbate molecule.
0) surface
0) facet is the most stable surface of hexagonal wurtzite ZnO,55–57 exposing three-coordinated Zn (denoted as Zn3c) and O (denoted as O3c) atoms (Fig. S1). Atoms in the subsurface and the bulk are four-coordinated and are denoted Zn4c and O4c, respectively. As shown in Fig. 1a, atoms of a single type (O or Zn) are linearly arranged along the [1
10] direction. Along the [0001] direction, Zn and O atoms alternate. Besides, there are two types of adjacent Zn–O pairs: one is closer and directly bonded, forming the Zn–O dimer, and the other is more distant, without direct bonding, resembling a trench, and is referred to as the Zn–O trench.58 Here, “dimer-Zn3c” denotes the Zn atom that pairs with a given O3c atom to form a Zn–O dimer, and “trench-O3c” denotes the O atom that pairs with a given Zn3c atom to form a Zn–O trench; “dimer-O3c” and “trench-Zn3c” are defined similarly.
For the O3c atoms on the ZnO (10
0) surface, the formation energy of a thermally induced OV is 3.39 eV (Fig. 1d), while hydrogen-induced OV formation is thermodynamically much more favourable, with a formation energy of only 0.33 eV. After heterolytic dissociation of H2 on a Zn–O pair, the H atom adsorbed on Zn3c migrates to the O3c site to form H2O. The subsequent desorption leaves an OV (Fig. S2). We define OVs generated via H species adsorbed on a Zn–O dimer as dimer-H induced OVs, and those from H adsorbed on a Zn–O trench as trench-H induced OVs. The rate-determining step (RDS) of OV formation is H2O formation rather than H2 dissociation. The energy barriers of the RDS are 1.55 eV and 2.05 eV for dimer-H and trench-H induced OV formation, respectively, indicating that OV formation is kinetically more favourable when H2 dissociates on a Zn–O dimer. In contrast, the removal of a subsurface O4c atom is more endothermic, with a formation energy of 3.84 eV (Fig. S3a). Our AIMD sampling indicates that H species are unlikely to adsorb onto Zn4c–O4c due to the higher endothermicity (Fig. S3b and c). Consequently, neither thermally nor hydrogen-induced vacancies are likely to originate from O4c.
ELF analysis (Fig. 1c) reveals charge localization at the OV site, which attracts nearby Zn ions—one on the surface (denoted as Zna) and two in the subsurface (denoted as Znb and Znc, which are symmetry-equivalent). The Zn–Zn distances decrease from 3.01 and 3.29 Å to 2.56 and 3.00 Å, approaching the bond lengths in hexagonally close-packed metallic Zn (2.66 and 2.94 Å, Fig. S1d). The Zna–Znb interaction is significantly strengthened upon OV formation, as indicated by an increase in the −ICOHP value from 0.02 eV to 0.50 eV (Fig. 1e and f). The three Zn atoms on the defect surface lose 0.62, 0.94, and 0.94 e, respectively. These values are smaller than the corresponding 1.16, 1.21, and 1.21 e on the pristine surface (Fig. 1a and b). This suggests a partial reduction of Zn and an enhanced metallic character. Together, these results indicate that OV formation induces the aggregation of surrounding Zn ions into a metallic Zn3 cluster-like structure, and herein, we refer to this structural motif as OV–Zn3.
The formation of OV–Zn3 suppresses H adsorption at the O3c site due to the more endothermic adsorption energy, accompanied by a decrease in charge transfer to the surface from 0.61 e (Fig. 2b) to 0.57 e (Fig. 2d). On the pristine ZnO (10
0) surface, the O3c atom directly binding the H atom gains only 0.08 e, and the Bader charges of the other atoms change only slightly. Thus, the excess charge is nearly uniformly delocalized over the entire slab, indicating that ZnO possesses a strong capacity for electron storage, consistent with the earlier work of Zhang et al.58 Conversely, H adsorption at the Zn3c site is promoted by the presence of OV–Zn3. When adsorbed at the OV site, the H atom nearly occupies the original O3c position, delocalizing the charge and stabilizing the defect structure (Fig. 2f). The H adsorbate can alternatively be viewed as occurring at the hollow site of the Zn3 cluster, which acquires significantly more charge from the surface by binding to multiple Zn atoms compared to its adsorption at the regular Zn3c or Zna site. Thus, H adsorption at the OV site is substantially stronger than that at other Zn sites, as reflected by its notably lower adsorption energy, making it possible for H to adsorb at the OV site under typical experimental conditions.
The formation of OV–Zn3 markedly alters the interaction between hydrogen and the ZnO (10
0) surface, especially when H is adsorbed on Zn. On the defect surface, the OV–Zn3 motif acts as an electron reservoir, characterized by its strong electron-donating ability and relatively weak electron-accepting capacity. Fig. 2e and S4d clearly reveal the disappearance of localized electron density at the OV site, confirming the role of the OV–Zn3 motif as a key source of electrons in the charge redistribution process. Changes in the localized charge of the OV–Zn3 region induced by H adsorption further leads to lattice distortion. H adsorption at an O3c site increases the local electron density at Zna and Znc, shortening the Zna–Znc and Znb–Znc distances while slightly increasing the Zna–Znb distance. Thus, the Zna–Znc bond is strengthened with the change of the −ICOHP value from 0.50 eV to 0.56 eV, whereas the Zna–Znb bond is weakened with the −ICOHP value changing from 0.50 eV to 0.45 eV (Fig. 3a). In contrast, when an H atom adsorbs on Zna (Fig. 3b), depletion of the localized electron density increases the electrostatic repulsion among the Zn cations, driving their further separation. Accordingly, the −ICOHP value between Zna and Znb (or Znc) decreases from 0.50 eV to 0.03 eV. When H adsorbs at the OV site (Fig. 3c), local electrons are likewise depleted. The Zn–Zn separations become shorter, and the system is stabilized because the H atom simultaneously coordinates to all three Zn atoms.
In addition to tuning the adsorption energies, H co-adsorption induces significant charge redistribution and affects interfacial bonding. On the defect ZnO (10
0) surface, upon the adsorption of an additional H atom at a dimer- or trench-O site, the pre-adsorbed H on Zn gains some extra electron density, 0.05 e and 0.08 e, respectively (Fig. 2e and 4c and d). Conversely, if the second H adsorbs at a dimer- or trench-Zn site, the pre-adsorbed H on O may lose some electron density, 0.00 e and 0.05 e (Fig. 2d and 4e and f), respectively. Thus, the co-adsorption results in an additional electron transfer from the O-bound H to the Zn-bound H, as visualized by charge-density difference maps (Fig. 4c–f). Compared to single H adsorption, co-adsorption of two H atoms on a Zn–O dimer of the defect surface affects the bond significantly. The H–O bond is weakened, as reflected by a decrease in the −ICOHP value from 3.83 to 3.76 eV (Fig. S5a and b). Conversely, the H–Zn bond is strengthened, with the −ICOHP increasing from 0.96 to 1.00 eV (Fig. S5c and d). Similar trends are also observed on the pristine surface (Fig. S6 and S7). Thus, co-adsorption of two H atoms facilitates additional electron transfer from H–O to H–Zn, weakening the former and strengthening the latter. In addition to electron redistribution, the electrostatic attraction between the oppositely charged H atoms further enhances the stability of the co-adsorption configuration.
When two H atoms are co-adsorbed on adjacent Zn–O pairs, the adsorption of a third H atom at a neighbouring site remains thermodynamically promoted (Fig. 4b), although this effect is notably weaker than that observed for single H pre-adsorption. For the defect surface, when two H atoms are co-adsorbed on a Zn–O dimer (Fig. S8), the adsorption energies of a third H atom at the adjacent O3c or Zn3c site along the [0001] direction are −0.36 eV and 1.36 eV, respectively. If the two H atoms are co-adsorbed on a Zn–O trench, the adsorption energies of a third H at the adjacent O and Zn sites (Fig. S9) along the [0001] direction are −0.57 eV and 0.92 eV, compared to the corresponding −0.21 eV on O and 1.43 eV on Zn for the clean surface. The pristine surface exhibits the same promotional effect (Fig. S10–S12), indicating that this phenomenon is independent of the presence of oxygen vacancies.
It is also evident that H atoms preferentially co-adsorb in pairs on the Zn–O dimer rather than on the Zn–O trench along the [0001] direction. As shown in Fig. 4a, the adsorption of a second H atom on a dimer site (O or Zn) is more exothermic than on a trench site. Furthermore, Fig. 4b and S10 reveal that when two H atoms are pre-adsorbed on a Zn–O trench, the adsorption of an additional H atom is even more exothermic than when they are pre-adsorbed on a dimer site. This is because co-adsorption on a Zn–O dimer results in isolated trench sites, while co-adsorption on a Zn–O trench leads to isolated dimer sites, which are more favourable for subsequent H adsorption.
Indeed, the next H atom preferentially adsorbs at a neighbouring site along the [0001] direction, rather than along the [1
10] direction. For the pristine surface, when a single H atom is adsorbed at an O or Zn site, the adsorption energies for a second H atom on the same type of site along the [1
10] direction are 0.19 eV and 1.79 eV, respectively, while these values are 0.28 eV and 0.55 eV for the defect surface (Fig. S13). Similarly, on the defect surface with a pre-adsorbed H pair on a Zn–O dimer, the adsorption energies for the third H atom on the adjacent Zn and O sites along the [1
10] direction are 1.69 eV and −0.11 eV, respectively, while for a pre-adsorbed H pair on a Zn–O trench, these values are 1.52 eV and −0.09 eV (Fig. 4b and S8–S10). These less favourable adsorption energies may be attributed to electrostatic repulsion between adjacent hydrogen atoms of the same polarity along the [1
10] direction.
0) surfaces are −0.48 eV and −0.42 eV, respectively, closely matching the experimental value of −0.52 eV.59 In contrast, chemisorption at the Zn–O trench sites is less exothermic, with values of −0.23 eV and −0.21 eV. However, the energy barriers of heterolytic dissociation are higher at the dimer sites, both at 0.52 eV, while those at the trench sites are lower, at 0.21 eV for the pristine surface and 0.19 eV for the defect surface (Fig. 1d and 5a). Thus, H2 activation is kinetically more favourable at the Zn–O trench, despite being thermodynamically more favourable for the adsorption at the Zn–O dimer. Due to the distinct nature of the adsorption sites, this does not contradict the Brønsted–Evans–Polanyi (BEP) relationship.60 Fig. S15 shows that H2 chemisorption at both dimer and trench sites does not induce charge delocalization at the vacancy, and the activation primarily involves the adsorption site and its nearby atoms. Energetic and electronic structure analyses indicate that H2 activation is nearly identical on the pristine and defect surfaces, which can be attributed to the stabilizing effect of H adsorption at the O3c sites, preserving the electronic and structural integrity of the OV–Zn3 motif. The possibility of H2 dissociative adsorption at the subsurface Zn4c and O4c sites is also studied. We estimated its energetics by AIMD sampling because the hypothetical dissociation state is highly unstable and cannot be effectively studied by routine structural optimization. Placing a pair of H atoms at appropriate initial distances on the Zn4c–O4c dimer site and on the Zn4c–O4c trench, the H atoms recombined into H2 after only 13 fs and 18 fs, with reaction energies of −3.55 eV and −4.69 eV (Fig. S3b and c), respectively. This indicates that dissociative adsorption of H2 at subsurface Zn4c and O4c sites is unlikely to occur under typical experimental conditions.
As discussed in the previous section, H atoms preferentially adsorb along the [0001] direction, leading to the experimentally observed 1D hydrogen chains.32,37 We simulated the growth of hydrogen chains by calculating the sequential chemisorption of two H2 molecules along the [0001] direction on ZnO (10
0) surface. As shown in Fig. 5a, the chemisorption of the second H2 at the Zn–O dimer is more exothermic (−0.73 eV and −0.72 eV for the pristine and defect surfaces) than the first, but the energy barriers are nearly unchanged at 0.51 eV and 0.49 eV, indicating no significant kinetic effect. In contrast, at the Zn–O trench, dissociation of the second H2 is even more exothermic (−0.90 eV and −0.87 eV for the pristine and defect surfaces) and essentially barrierless (0 and 0.02 eV), suggesting a highly favourable pathway. The significant reduction in the energy barrier is primarily attributed to the isolated dimer-O and dimer-Zn sites formed by H2 chemisorption at the Zn–O trench (Fig. 5b and S16a and S17a). The isolated dimer O and Zn sites provide stronger binding sites for H atoms from subsequent H2 dissociation, thereby facilitating continuous H2 activation and hydrogen chain growth. In contrast, isolated trench-O and trench-Zn sites (Fig. S18a and S19a) formed by H2 adsorption at the Zn–O dimer have much lower affinity for additional H atoms and do not markedly promote sequential H2 dissociation, consistent with the discussion in the previous section. Besides, the dissociative adsorption of a second H2 molecule along the [1
10] direction is more difficult (Fig. S16–S20), indicating that hydrogen chain growth along this direction is energetically and kinetically disfavoured. Therefore, hydrogen chains grow continuously only along the [0001] direction but are discontinuous along the [1
10] direction, resulting in their characteristic one-dimensional morphology.
We note that diffusion may occur due to the varying adsorption strengths of H at different sites. In fact, the H atoms adsorbed in pairs on the Zn3c–O3c site are very stable. When two H atoms are co-adsorbed on the Zn3c–O3c dimer, the migration of the H atom on Zn3c to a neighbouring trench-O3c site (Fig. S21c) is endothermic by 0.26 eV, with a significant energy barrier of 1.17 eV. Similarly, when two H atoms are co-adsorbed on the Zn3c–O3c trench, the migration of the H atom on Zn3c to a neighbouring dimer-O3c site (Fig. S21c) is also slightly endothermic by 0.06 eV, with an energy barrier of 1.13 eV (Fig. S22). These results demonstrate that when an H atom is co-adsorbed at an adjacent O site (dimer-O or trench-O), its migration from neighbouring Zn3c becomes thermodynamically and kinetically more difficult, which can be attributed to the cooperative effect of co-adsorption as mentioned earlier, resulting in the 1D hydrogen chains. This is different from the conclusion of Song et al., who believed that even if H atoms are adsorbed in pairs, H on Zn3c could easily migrate to adjacent O3c atoms.34 In fact, early infrared studies of H2/D2 chemisorption on predominantly (10
0)-terminated ZnO already revealed type-I hydrogen species, where the Zn–H stretching modes appear along with the characteristic O–H bands, attributed to heterolytic H2 dissociation at the Zn–O pairs.61 Further high-resolution FTIR “spectral-ratio” measurements identified weak absorptions near 840 and 820 cm−1 attributable to coupled δ(OH) and δ(ZnH) bending modes,62 while incoherent inelastic neutron-scattering studies identified Zn–H bending and stretching modes around 829 and 1708 cm−1, in excellent agreement with the IR experiment.63 More recent DRIFTS/INS studies64 and theoretical analyses have further confirmed this picture, consistently treating type-I Zn–H hydrides associated with Zn–O pairs as experimentally detectable but minor H species on ZnO surfaces. This also shows that H species on Zn3c do not completely migrate to O3c atoms.
In addition to the regular Zn3c–O3c sites, H2 can also be activated at OV–Zn3. We examined seven possible dissociation pathways involving this site (Fig. S23, Table S1). Unlike H2 activation at the Zn3c–O3c sites, where the vacancy plays a negligible role, here the OV–Zn3 motif directly mediates H2 activation (Table S2). Among these pathways, the heterolytic dissociation of H2 across Zna and O3c along the [0001] direction is endothermic by 0.17 eV but exhibits the lowest energy barrier of 0.60 eV, rendering it kinetically the most favourable (Table S1 and Fig. S24). Additionally, H2 dissociation between the OV and Zna sites is slightly exothermic by −0.10 eV, yielding the thermodynamically most stable configuration, but involving a high energy barrier of 1.42 eV. Another pathway, involving H adsorption at the Zna–Znb and Zna–Znc bridge sites—resembling homolytic H2 dissociation on metal surfaces—is both kinetically (2.79 eV) and thermodynamically (1.39 eV endothermic) unfavourable. ELF and Bader charge analyses (Fig. S25) confirm the homolytic nature of both processes. The remaining four pathways are all exothermic, with energy barriers exceeding 1.00 eV. These results show that H2 activation at OV–Zn3 is clearly more difficult than that at Zn3c–O3c, which can proceed even at temperatures as low as 20 K,32 but that at OV–Zn3 requires high temperatures. Since reactions such as CO2 hydrogenation to methanol typically occur at 473–573 K,65–69 this makes H2 activation at OV–Zn3 possible under typical experimental conditions for these reactions.
0) surface, followed by H diffusion. Consequently, the surface coverage of hydrogen species can vary with the reaction conditions. Based on our studies on H2 activation and diffusion, we rationalize the experimental observation of Song et al.34 that H species can be clearly observed at OVs at 200 °C, whereas further increasing the temperature to 300 °C leads to a decrease in H coverage at OVs.
At a low temperature, the OV concentration on the ZnO (10
0) surface is negligible, and the 1D hydrogen chains observed at ∼20 K mainly32 originate from sequential H2 chemisorption at adjacent Zn–O trenches. Upon heating to 423 K, low concentrations of OVs are formed. However, due to the relatively high energy barrier for H2 dissociation at the OV–Zn3 sites, OV-mediated H2 activation is unfavourable at a low OV concentration. Once the temperature reaches ∼473 K, homolytic H2 dissociation at OV–Zn3 to form H atoms adsorbed at the OV and Zna sites becomes kinetically favourable, considering the magnitude of its energy barrier of 1.43 eV. This leads to the emergence of the characteristic 1H MAS NMR resonance at ∼8.4 ppm assigned to hydride species located at OVs, accompanied by a concurrent decrease in the EPR intensity and the O 1s XPS signal associated with paramagnetic OVs.34
The co-adsorption of two H atoms at the OV and Zna sites remains relatively stable within a certain temperature range. As shown in Fig. 6, we examined two possible pathways for the transformation of this co-adsorption configuration into a thermodynamically more favourable state, namely, a pair of H atoms adsorbed on a Zn–O dimer. In pathway I (Fig. 6c), H on Zna migrates along the [1
10] direction to Zn3c with an energy barrier of 0.86 eV, followed by the migration of another H at OV to O3c with a much higher energy barrier of 1.57 eV as the rate-determining step. In pathway II (Fig. S26), the H atom at the OV site first migrates to the O3c atom with an energy barrier of 2.10 eV, followed by H migration from Zna to Zn3c with an energy barrier of 0.25 eV. Clearly, pathway I is kinetically more favourable, although the energy barrier of its RDS still exceeds that of H2 dissociation at the OV–Zna site. Besides, the intermediate state is less stable than the initial co-adsorption state by 0.57 eV. Thus, the forward process is unlikely to occur unless at elevated temperatures, which enable the two H atoms to sequentially diffuse along the [12
0] direction to the Zn–O dimer and form the thermodynamically more stable configuration. Within a temperature window, H species can accumulate at the OV–Zn3 motif rather than fully migrating to the O3c–Zn3c dimers. In addition, H species adsorbed at the OV–Zn3 motif are unlikely to recombine into H2 and desorb, because the recombined state is thermodynamically less stable than the co-adsorbed state. Upon further increasing the temperature, more H species initially on OV–Zn3 can diffuse to the O3c–Zn3c sites via pathway I, with an energy barrier only 0.14 eV higher than that of H2 dissociation at the OV–Zn3 site. We note that Song et al. further showed that at 573 K (300 °C), the 1H MAS NMR signal at ∼8.4 ppm, attributed to H species at Ovs, becomes weaker than the signal observed after H2 treatment at 473 K (200 °C).34 Our microkinetic simulations based on the above analysis successfully reproduce this behaviour and explicitly reveal the decrease in H coverage at the OV sites (Fig. 6b). Thus, the outward migration mechanism of H atoms at the OV–Zn3 motif proposed here can explain the experimentally observed decrease, in addition to the bulk O replenishment of surface OV sites proposed by Song et al.34
H2 chemisorption onto the two bridge sites of Zn3 (i.e., Zna–Znb and Zna–Znc) is unlikely to occur under typical reaction conditions, due to the very high energy barrier of 2.79 eV. Even if it occurs at much higher temperatures, the resulting species will spontaneously convert into a more stable configuration without any kinetic barrier, where the two H atoms are co-adsorbed at the OV and Zna sites. These H atoms can subsequently diffuse along pathway I to a Zn–O dimer, where they can recombine and desorb as H2. We also considered H diffusion from the OV and Zna sites along the [0001] direction (Fig. S27). Although the RDS has a much lower energy barrier of 1.30 eV, the final states are thermodynamically unstable, making the reverse process both kinetically and thermodynamically more favourable (Fig. S28). Therefore, the observed decrease in the H coverage at the OV sites should be primarily attributed to the diffusion of H adsorbates along the [1
10] direction. In addition to the previous reports on the migration of H atoms from Zn3c to O4c, we also considered the possible migration of H species from Zna to O4c. Our simulations show that upon H adsorption on O4c, it pulls O4c from the subsurface to the surface. This was also observed in the “heterolytic-1 + migration-1 pathways” reported in the previous study.34 Our calculations show that the energy barrier for the direct migration of H species from Zna to Zn3c is lower than that to O4c, because the former does not involve the high energy required for pulling O4c from the deep subsurface to the surface (Fig. S29).
0) surface were investigated by DFT calculations. Our calculations show that H2 adsorption at the Zn–O trench sites is the most favourable for the formation of the OV–Zn3 motif, which possesses strong electron-donating ability and relatively weak electron-storage capacity. In addition, H adsorbates at the OV or O site stabilize the defect structure, whereas those on the Zn site induce electron transfer from the OV, leading to an increased Zn3 separation. Pre-adsorbed H species can accelerate H2 dissociation at neighbouring sites, and H atoms preferentially align along the [0001] direction, yielding the experimentally observed 1D hydrogen chains. Our DFT calculations indicate that H2 dissociation at the OV–Zn3 motif and the accumulation of H species at the OVs become feasible only at elevated temperatures. However, at higher temperatures, migration of these H species toward the Zn3c–O3c dimers leads to a gradual decrease in the H coverage at OVs. Our theoretical insights are consistent with the experimental observations and are further confirmed by our microkinetic simulations.
These simulations highlight the critical role played by the OV–Zn3 motif in governing the interaction between the hydrogen species and the surface. The insights gained from this study offer an important perspective for the rational design of ZnO-based hydrogenation catalysts, by leveraging defect engineering and thermal control to modulate lattice distortion, hydrogen activation sites, and the directional spillover of hydrogen species.
Supplementary information is available. See DOI: https://doi.org/10.1039/d5cy01291k.
| This journal is © The Royal Society of Chemistry 2026 |