Synergistic effects of oxygen vacancies and Pd single atoms on Pd@TiO2−x for efficient HER catalysis

Xiaoyu Zhao Kai Gao Sen Xue Wei Ran Rui Liu

Citation:  Xiaoyu Zhao, Kai Gao, Sen Xue, Wei Ran, Rui Liu. Synergistic effects of oxygen vacancies and Pd single atoms on Pd@TiO2−x for efficient HER catalysis[J]. Chinese Chemical Letters, 2025, 36(6): 110309. doi: 10.1016/j.cclet.2024.110309 shu

Synergistic effects of oxygen vacancies and Pd single atoms on Pd@TiO2−x for efficient HER catalysis

English

  • The widespread use of fossil fuels has led to severe energy consumption and environmental pollution [1,2]. Hydrogen energy, due to its high energy density, low cost, lack of pollution, and zero carbon emissions, is considered the most promising clean energy source [3]. Electrochemical water splitting driven by renewable electricity is an ideal and efficient hydrogen production technology. However, the sluggish kinetics of the hydrogen evolution reaction (HER) at the cathode significantly hinders large-scale hydrogen production [4-6]. To enhance the kinetics of the HER, reduce the reaction overpotential, and design efficient and durable HER electrocatalysts is crucial. Currently, Pt@C and Pd@C catalysts are the most efficient commercial hydrogen evolution electrocatalysts [7]. However, their high cost and material instability, prone to deactivation due to corrosion during usage, hinder their widespread application. Therefore, there is an urgent need to explore more economically efficient electrocatalysts [8-10].

    Considerable efforts have been devoted to developing novel precious metal HER electrocatalysts (e.g., Ru, Rh, Pt, and Ir) [11-13]. Among them, Pd is considered one of the most promising candidate materials [14]. Numerous Pd-based catalyst materials have been applied to electrocatalytic hydrogen evolution. Heydari-Bafrooei et al. synthesized a Pd/WS2 catalyst with inorganic fullerene WS2-supported Pd nanoparticles using a sol-gel immobilization method [15]. By encapsulating Pd nanoparticles within WS2 cages for the HER reaction, the catalyst material’s activity was improved, demonstrating excellent long-term stability [16]. Liu et al. through electrodeposition, anchored Pd nanoparticles onto MoSx/TiO2 nanotube arrays, enhancing the carrier’s electrical conductivity [17]. The introduction of Pd nanoparticles significantly improved the carrier’s conductivity, and the electronic structure of MoSx weakened the Pd-H bond, promoting hydrogen desorption and enhancing electrocatalytic activity [18]. Zheng et al. used a highly conductive MOF as a template to prepare a Pd@MOF-74-M composite catalyst [19]. The small size and high dispersion of Pd nanoclusters, as well as the effective protection provided by the MOF framework, exhibited excellent HER catalytic activity [20]. Gao et al. successfully synthesized sub-nanometer Pd clusters in porous cerium dioxide (Pd NCs@CeO2) [21]. The electron transfer between Pd and O atoms in the material effectively regulated the adsorption state of active hydrogen on Pd clusters, promoting the HER catalytic activity of the composite system [22]. To further reduce the cost of precious metals and maximize atomic utilization, single-atom catalysis has garnered significant attention [23]. Single-atom catalysts (SACs), composed of single metal atoms fixed on carrier materials, can maximize the exposure of active sites, improve atomic utilization, reaction selectivity, and enhance the system’s electrocatalytic activity [24]. Yu et al. reported graphene (GDY)-supported zero-valent Pd atoms on ultrathin nanosheets for direct application as a three-dimensional flexible hydrogen evolution cathode [25]. The successful anchoring of individual dispersed palladium atoms on GDY maximized the atomic efficiency and the number of active sites, greatly improving catalytic activity. Zhu et al. anchored isolated Pd atoms on Ti3C2Tx MXene, optimizing hydrogen adsorption on MXene and effectively enhancing water molecule dissociation through strong electronic interactions between MXene layers, improving HER activity [26]. However, the low electrical conductivity, low loading density, and poor stability and tolerance of single-atom catalysts severely limit their application and catalytic efficiency [27]. Therefore, selecting a suitable stable material as a precious metal carrier is necessary.

    To enhance the catalytic performance, stability, and selectivity of precious metals, researchers have developed various carriers, such as metal oxides, metal nitrides, metal phosphides, and metal sulfides [28-32]. For example, Wang et al. coated ruthenium with ruthenium oxide on TiO2, demonstrating excellent stability for hydrogen electro-oxidation [33]. Zeng et al. reported a highly efficient nano-composite electrocatalyst composed of Pd NPs uniformly embedded in mesoporous oxygen-deficient TiO2 nanospheres [34]. Doping Ti4+ sites induced the formation of abundant oxygen vacancies in mesoporous TiO2 nanospheres, enhancing the activity and stability of the composite material. Due to its richness and good physicochemical properties, titanium dioxide has been widely studied as a substrate for depositing precious metals to enhance catalytic performance and stability [35]. Compared to TiO2, TiO2−x with abundant oxygen vacancies has received more attention due to its superior conductivity, catalytic activity, and stronger stability [36,37]. Zhang et al. reported that TiO2−x substrates not only stabilized MoS2 but also promoted the adsorption and dissociation of water molecules during the HER process [38]. The composite material MoS2/TiO2−x@Ti exhibited excellent catalytic performance. Additionally, Zhang et al. designed a structure decorating TiO2−x nanotube arrays with CoP3/CoMoP nanosheets [39]. The oxygen vacancy-rich TiO2−x nanotube structure provided more active sites, enhancing HER performance. Wu et al. developed a one-step rapid method without the need for high temperature and inert gas protection to synthesize oxygen-deficient black TiO2-supported sub-nanometer Pt clusters with metal-support interactions [40]. The metal-support interactions between Pt and black TiO2 endowed the prepared composite material with excellent electrocatalytic performance for HER [41]. The coupling effect between oxygen vacancies and Pt optimized the electron density around Pt, thereby optimizing the adsorption free energy of hydrogen and improving the electrocatalytic performance of HER in acidic media [42]. Moreover, the strong interaction between Pt and oxygen-deficient titanium dioxide significantly enhanced the stability of the designed electrocatalyst [43].

    In this study, TiO2(B) was selected as the precursor material. Through annealing in an argon atmosphere, it was transformed into oxygen-deficient TiO2−x as the carrier. Using a photo-reduction deposition method, Pd single atoms were successfully anchored on the surface of oxygen-deficient titanium dioxide. The newly synthesized Pd@TiO2−x(1.5%) composite material (TiO2−x material with Pd content of 1.5%) exhibited significantly superior HER activity. Benefiting from the ultra-stable oxygen-deficient titanium dioxide carrier, the composite material demonstrated robust long-term electrocatalytic stability, with almost no decline in activity after 1000 cycles. Further density functional theory (DFT) theoretical calculations indicated that the presence of oxygen vacancies improved the stability and catalytic activity of the material, promoting the generation of more active sites and enhancing the adsorption of active hydrogen by precious metal Pd. Additionally, by adjusting the Pd loading amount, Pd-loaded TiO2−x materials of different morphologies were prepared. The presence of single-atom Pd significantly increased the abundance of oxygen vacancies, and Pd, as the main active center, greatly promoted the generation of active hydrogen intermediates, reducing the Gibbs free energy of hydrogen and accelerating the electrocatalytic kinetics of HER. The synergistic effect of single atoms and oxygen vacancies collectively promoted the high catalytic activity and stability of the composite material.

    Transmission electron microscopy (TEM), and elemental mapping analysis were performed using a Philips Tecnai F30 at 300 kV. Powder X-ray diffraction (XRD) was performed on Panalytical X′pert PRO X-ray diffractometer with monochromated Cu Kα radiation (40 kV, 30 mA). Inductively coupled plasma atomic mass spectrometry (ICP-MS) measurements were performed on PerkinElmer. X-ray photoelectron spectroscopy (XPS) was carried out on PHI5000VPIII hemispherical electron energy analyzer with monochromatized Al Kα radiation (1486.6 eV). Electron paramagnetic resonance (EPR) was performed at Bruker EMX-10/12 with the temperature at 293 K.

    To prepare the catalyst dispersion, Pd@TiO2−x material (10 mg) and XC-72R carbon black (5 mg) were placed in a mixture of isopropanol (0.5 mL) and 5% Nafion (0.05 mL). The mixture was sonicated for 1 h, and then 10 µL of the dispersion was drop-cast onto a glassy carbon electrode, resulting in a loading rate of 0.013 mg Pd/cm2 for the catalyst. The addition of carbon black to the catalyst dispersion enhances the electronic conductivity of the electrode surface and the dispersibility of the catalyst. The electrode preparation process for other materials was identical, using commercial Pd@C (10%) and Pt@C (10%) as reference materials.

    All electrochemical tests were conducted using a three-electrode system on a CHI440C electrochemical workstation, with a Pt mesh (20 × 20 mm) as the counter electrode, Ag/AgCl as the reference electrode, and glassy carbon as the working electrode. All tests were performed in a nitrogen-saturated 0.5 mol/L H2SO4 solution. Potentials were referenced to the reversible hydrogen electrode (RHE), and the results were calculated using Eq. 1.

    (1)

    Linear sweep voltammetry (LSV) was performed at a scan rate of 5 mV/s after 30 cyclic voltammetry (CV) cycles of preconditioning. Tafel slope was determined from the LSV curve. Catalyst stability was assessed by performing 1000 CV cycles (100 mV/s). Chronoamperometry was conducted by recording the time-current density curve at 10 mA/cm2. Double-layer capacitance (Cdl) was determined from CV curves at various scan rates (20, 40, 60, 80, 100 mV/s) within a specific potential window.

    DFT was employed as implemented in the Vienna Ab initio simulation package (VASP) for all calculations. The exchange-correlation potential was described using the generalized gradient approximation of Perdew-Burke-Ernzerhof (GGA-PBE). The projector augmented-wave (PAW) method was utilized to treat interactions between ion cores and valence electrons. The plane-wave cutoff energy was fixed at 500 eV [44]. The structural models were relaxed until the Hellmann-Feynman forces were smaller than −0.02 eV/Å and the change in energy was smaller than 10−5 eV.

    During the relaxation, the Brillouin zone was represented by a Γ centered k-point grid of 6 × 6 × 1 [45]. Grimme’s DFT-D3 methodology was employed to describe the dispersion interactions among all the atoms in the adsorption models.

    According to XRD analysis, the main component of TiO2−x is rutile, thus the model structure of rutile is used as the carrier structure [46]. The TiO2 (101) surface was modeled using two three-layer periodic slabs. The lattice constants for the (3 × 2) TiO2 (101) unit cell were set as follows: a = 11.05 Å, b = 11.49 Å, c = 25.00 Å. To further investigate the role of oxygen vacancies, a single oxygen was removed from the (3 × 2) TiO2 (101) surface, according to reference [47]. Simultaneously, to account for the influence of the surface periodic structure, a vacuum layer was set to 15.00 Å.

    Single Pd, optimized Pd13 clusters, and Pd(111) species were adsorbed on the TiO2 (101) surface with a single oxygen removed, representing single-atom Pd, cluster Pd, and nano Pd, respectively, for the structural optimization of Pd-based catalysts [48]. The optimized configurations were subjected to Bader charge, differential charge, and density of states analyses [49]. Multiple adsorption configurations and positions were considered for all adsorbate species, and the subsequent calculations selected the lowest energy and most stable configurations. The formula for calculating hydrogen adsorption energy (ΔEH) is as follows:

    (2)

    EM+H represents the total energy associated with the absorption of a single hydrogen atom on the surface of Pd@TiO2−x material [50]. EM denotes the total energy of Pd@TiO2−x, and EH2 signifies the energy of a gaseous hydrogen molecule (approximately −6.7 eV in this study) [51]. The Gibbs free energy correction for hydrogen absorption incorporates adjustments for entropy and zero-point energy:

    (3)

    ΔEH represents the adsorption energy of hydrogen, ΔEZPE denotes the difference in zero-point energy between the adsorbed state of hydrogen and gaseous hydrogen, and ΔSH signifies the entropy difference between the adsorbed state of hydrogen and gaseous hydrogen [52]. Since the contribution of vibrational entropy for the adsorbed state of hydrogen is negligible, it can be ignored. The adsorption entropy for hydrogen gas is denoted as , where SH2 represents the entropy of the H2 gaseous phase under standard conditions. Therefore, the corrected Gibbs free energy is expressed as follows:

    (4)

    Fig. S1a (Supporting information) illustrates the precursor as the typical TiO2(B) phase. The Raman spectrum of TiO2(B), depicted in Fig. S1b (Supporting information), reveals Raman vibrational peaks at 256, 382, 422, 477, 553, 632 and 828 cm−1. These peaks confirm that the prepared precursor material is purely TiO2(B). Fig. 1a presents the XRD patterns of Pd-loaded TiO2−x with different proportions. Two peaks at 25.3° and 48.0° are attributed to the (101) and (200) crystal planes of anatase, respectively, and two peaks at 27.4° and 36.1° are assigned to the (110) and (101) crystal planes of rutile (JCPDS, 99-0008 and 99-0090, respectively) [53]. It can be inferred that, under argon atmosphere, after calcination at 800 ℃ for 3 h, TiO2(B) transforms into anatase and rutile phases, indicating a structural change. No distinct Pd characteristic diffraction peaks are observed in the XRD patterns of Pd@TiO2−x(1.5%) and Pd@TiO2−x(3%), attributed to the small size and good dispersion of single-atom Pd and Pd clusters, making them undetectable by XRD. In the XRD patterns of Pd@TiO2−x(6%) and Microscopy (TEM). The morphology of TiO2, as shown in Fig. S2 (Supporting information), appears as a two-dimensional (2D) nanosheet. The morphology of TiO2−x, as shown in Fig. 1b, is closer to a particulate form. Figs. 1c and d are TEM images of TiO2−x further magnified after increasing the resolution, where lattice fringes with spacings of 0.351 nm and 0.324 nm are measured [54]. These spacings correspond to the (101) lattice plane of the TiO2 anatase phase and the (110) lattice plane of the rutile phase, respectively, consistent with the XRD results mentioned earlier. The morphology of Pd@TiO2−x(1.5%) and Pd@TiO2−x(3%) is shown in Figs. 1e and f, where no obvious Pd aggregation is observed, and there are no distinct Pd lattice fringes. This is attributed to the low Pd loading in the 1.5% material, and the extremely small size of Pd single atoms, making them challenging to observe clearly by TEM. As shown in Fig. 1f, Pd@TiO2−x(3%) predominantly exhibits Pd clusters in an entirely exposed form. This confirms the successful anchoring of Pd clusters on the TiO2−x material. The morphology of Pd@TiO2−x(6%) and Pd@TiO2−x(10%) is shown in Figs. 1g and h, indicating uniform distribution of Pd NPs on the carrier. Furthermore, by increasing the resolution and measuring lattice fringe spacings, distinct lattice fringes with a spacing of 0.224 nm corresponding to the (111) crystal plane of Pd NPs are clearly observed. Particle size distribution analysis shows that the nanoparticle size of Pd@TiO2−x(6%) is approximately 5.91 nm, and that of Pd@TiO2−x(10%) is approximately 9.96 nm, as shown in Figs. S3a and b (Supporting information). To verify the existence of single atoms in Pd@TiO2−x(1.5%), aberration-corrected high-angle annular dark-field scanning transmission electron microscopy (AC-HAADF- STEM) was employed for characterization, as illustrated in Fig. 1i. This clearly shows that Pd is predominantly distributed as individual atoms uniformly on the surface of the TiO2−x carrier, confirming the successful preparation of single atoms (Figs. S4a and b in Supporting information).

    Figure 1

    Figure 1.  (a) XRD patterns of TiO2-x, Pd@TiO2-x(1.5%), Pd@TiO2-x(3%), Pd@TiO2-x(6%), Pd@TiO2-x(10%). (b) TEM images of TiO2−x. (c, d) TEM images of different crystal forms of TiO2−x. (e-h) TEM images of Pd@TiO2−x(1.5%), Pd@TiO2−x(3%), Pd@TiO2−x(6%), Pd@TiO2−x(10%). (i) HAADF-STEM image of Pd@TiO2−x(1.5%).

    The near-surface charge states of the catalyst materials were investigated using XPS. Figs. 2a and b present the XPS spectra of TiO2−x, revealing the predominant presence of Ti and O elements of TiO2−x. The Ti 2p peaks of TiO2−x are located at 464.7 eV and 459.0 eV, corresponding to the 2p1/2 and 2p3/2 states of Ti4+, respectively. These binding energies are lower than the Ti 2p peaks in pure anatase, which are at 465.5 eV and 459.5 eV, indicating a shift towards lower binding energy possibly due to electron gain from oxygen-deficient atoms in TiO2. In the O 1s spectrum, the three peaks at 529.7, 532.6, and 533.4 eV correspond to lattice oxygen in O-Ti, oxygen atoms near oxygen vacancies, and surface-adsorbed water molecules, respectively. Notably, the peak at 532.6 eV has a larger area, suggesting a significant presence of oxygen vacancies [55]. This is advantageous for fine-tuning the chemical and electronic structures of the material surface, promoting enhanced reactivity. To further validate the generation of oxygen vacancies, Raman spectroscopy was performed on the materials, as shown in Fig. 2c. The Raman peaks corresponding to the anatase and rutile phases of titanium dioxide were detected in TiO2−x, consistent with XRD results. The peak at 142 cm−1 is attributed to the Eg vibrational mode of the anatase phase, the peak at 515 cm−1 to the A1g + B1g vibrational mode of the anatase phase, the peak at 448 cm−1 to the Eg vibrational mode of the rutile phase, and the peak at 612 cm−1 to the A1g vibrational mode of the rutile phase. Anatase phase of titanium dioxide belongs to the tetragonal crystal phase, and in the tetragonal crystal system, changes in the vibrational peak of the Eg mode can serve as direct evidence of oxygen vacancy defect formation. The generation of oxygen vacancies can enhance the sample’s conductivity and charge transfer capability [56]. By comparing the most intense vibrational peaks of the Eg mode for all materials in the anatase phase (~142 cm−1), it is observed that the TiO2−x material is anchored by Pd atoms, lattice defects are induced by the strong Pd-O bonding interaction, resulting in more oxygen vacancies, and the crystallinity of TiO2−x changes, resulting in wider distribution of vibration modes [57,58]. Additionally, electron paramagnetic resonance spectroscopy (EPR) was employed to further confirm the presence of oxygen vacancies in the materials [59]. The results, as shown in Fig. 2d, reveal no apparent EPR peaks in TiO2, while TiO2−x and Pd@TiO2−x(1.5%) both exhibit a sharp peak at g = 2.003, consistent with the characteristic EPR peak of oxygen vacancies [60]. Moreover, at g = 1.983, there is no distinct Ti3+ peak, indicating that the oxygen vacancies in TiO2−x and Pd@TiO2−x(1.5%) are generated during the crystalline phase transition in a high-temperature argon atmosphere. Similar phenomena have been reported in previous literature. From the spectra, it is evident that the signal peak of oxygen vacancies is significantly enhanced when single-atom Pd is loaded onto the TiO2−x carrier [61]. This enhancement is attributed to the strong interaction between Pd and the TiO2−x carrier after the introduction of single-atom Pd, forming a robust Pd-O bonding, thereby facilitating the generation of oxygen vacancies on the carrier. By comparing the intensities of characteristic peaks, it is inferred that single-atom Pd, when combined with the TiO2 carrier, provides the maximum abundance of oxygen vacancies [62]. High-resolution X-ray photoelectron spectroscopy was performed on Pd-loaded TiO2−x with different proportions to determine the chemical states of different Pd forms, as shown in Fig. S5 (Supporting information). The Pd 3d5/2 peaks of Pd@TiO2−x(1.5%) (336.4 eV), Pd@TiO2−x(3%) (336.2 eV), Pd@TiO2−x(6%) (335.4 eV), and Pd@TiO2−x(10%) (335.1 eV) are located between Pd0 (335.0 eV) and Pd2+ (337.9 eV), indicating that Pd is in an oxidized state. With increasing Pd loading, the binding energy of Pd 3d shifts towards lower energy, suggesting a strong interaction between highly dispersed Pd single atoms and clusters with the TiO2−x carrier. The binding energies of Pd@TiO2−x(1.5%) and Pd@TiO2−x(3%) for Pd 3d3/2 are 341.7 eV and 340.9 eV, respectively. No peaks at 335.0 eV are observed for the single-atom and cluster forms, indicating the absence of distinct Pd-Pd bonds on these two materials [63]. This is consistent with TEM results, confirming the successful anchoring of isolated single atoms and clusters on the TiO2−x carrier. Furthermore, the higher binding energy of single-atom Pd compared to cluster Pd and nano Pd indicates a stronger interaction between single-atom Pd and the TiO2−x carrier.

    Figure 2

    Figure 2.  (a, b) High-resolution XPS Ti 2p and O 1s spectrum of TiO2−x. (c) Raman spectrums of TiO2−x, Pd@TiO2−x(1.5%), Pd@TiO2−x(3%), Pd@TiO2−x(6%), Pd@TiO2−x(10%). (d) EPR spectra of different catalyst.

    The hydrogen evolution reaction (HER) performances of TiO2, TiO2−x, Pd@TiO2−x(1.5%), Pd@TiO2−x(3%), Pd@TiO2−x(6%), Pd@TiO2−x(10%), Pd@C(10%), and Pt@C(10%) were systematically evaluated in a 0.5 mol/L H2SO4 solution saturated with N2 using steady-state linear sweep voltammetry (LSV). As depicted in Figs. 3a and b, TiO2 and TiO2−x exhibit large overpotentials of 256.5 mV and 193.5 mV, respectively. The presence of oxygen vacancies reduces the overpotential by approximately 25%, demonstrating that oxygen vacancies on the carrier surface contribute to an enhancement in catalytic activity.

    Figure 3

    Figure 3.  (a, b) HER polarization curves and overpotential histogram of as-prepared samples in 50 mmol/L H2SO4, respectively. (c) Schematic illustration of the HER mechanism on Pd@TiO2−x(1.5%). (d, e) Tafel plots and Slope histogram of prepared samples in 50 mmol/L H2SO4, respectively. (f) TOF curves with respect to Pd or Pt atoms and mass activities (based on Pd or Pt mass loading) at η = 50 mV of catalysts. (g) Corresponding current density scan rate curves to estimate the Cdl, respectively. (h) Cycling stability tests of Pd@TiO2−x(1.5%) by 1000 cycles. (i) Long-term chronoamperometry measurements at a current density of 10 mA/cm2 of Pd@TiO2−x(1.5%).

    This is attributed to the increased crystallinity and conductivity associated with the introduction of oxygen vacancies, leading to reduced charge transfer resistance and improved catalytic performance. Upon Pd loading onto the surface of TiO2−x, the overpotential decreases by about 90%, confirming that the high HER activity primarily originates from Pd species. Additionally, the overall electrocatalytic performance of Pd@TiO2−x surpasses that of commercial Pt@C (30 mV) and commercial Pd@C (51 mV), exhibiting higher HER activity. Remarkably, Pd@TiO2−x(1.5%) achieves a current density of 10 mA/cm2 at an overpotential of 24 mV with a low loading rate of 0.013 mg Pd/cm2. The strong Pd-O bonding interaction between single-atom Pd and the oxygen vacancies, this may be due to the introduction of heavier Pd atoms that affect the coordination structure of TiO2 [64], on the TiO2−x carrier facilitates the generation of oxygen vacancies, as confirmed by the EPR results [65]. The robust Pd-O interaction and increased oxygen vacancies synergistically contribute to the enhanced catalytic performance, as illustrated in Fig. 3c.

    Figs. 3d and e present Tafel test results for different materials. Pure TiO2 exhibits the highest slope, with a value of 161.72 mV/dec. The presence of oxygen vacancies in TiO2−x enhances the catalytic performance, resulting in a slope of 127.15 mV/dec. Upon loading precious metal Pd onto TiO2−x, the prepared Pd@TiO2−x materials exhibit a significant reduction in Tafel slope. The Tafel slopes for Pd@TiO2−x materials with different Pd contents are 41.9, 45.1, 36.7 and 35.9 mV/dec, indicating that the HER process of Pd@TiO2−x materials follows the Volmer-Heyrovsky mechanism, and materials with loaded Pd metals show faster reaction kinetics.

    The intrinsic activity of the catalysts was verified by turnover frequency (TOF), assuming that all Pd sites on the catalyst contacted the electrolyte and participated in the HER process. As shown in Fig. 3f, the TOF values of Pd@TiO2−x(1.5%) are 2.4 times, 3.5 times, 5.6 times, and 12.5 times higher than those of Pd@TiO2−x(3%), Pd@TiO2−x(6%), Pd@TiO2−x(10%), and Pd@C, respectively. When normalized to Pd loading, Pd@TiO2−x(1.5%) exhibits outstanding mass activity (2.63 A/mg) at an overpotential of 0.05 V, surpassing the mass activities of Pd@TiO2−x(3%), Pd@TiO2−x(6%), Pd@TiO2−x(10%), and Pt@C. This indicates that the loading of single-atom Pd substantially enhances the intrinsic activity of each catalytic site on the titanium dioxide surface. Notably, the catalytic activity of Pd single atoms exceeds that of Pd clusters and Pd nanoparticles.

    Fig. 3g displays the electrochemically active surface area of the catalyst materials. Pd@TiO2−x(1.5%) exhibits the highest double-layer capacitance value, significantly exceeding those of Pd@TiO2−x(3%), Pd@TiO2−x(6%), and Pd@TiO2−x(10%), indicating that single-atom palladium enlarges the active surface area of TiO2−x. Stability is a crucial factor in evaluating catalysts. Fig. 3h demonstrates that Pd@TiO2−x(1.5%) only attenuates by 1% after 1000 cycle scanning of LSV, maintaining good cyclic performance and verifying its stability in terms of catalytic performance. Furthermore, Fig. 3i shows Pd@TiO2−x(1.5%) at the current density of 10mA/cm2 was tested by chronopotentiometric method, and it was found that it only decreased by 0.8% after 10 h, which verified its outstanding stability. Fig. S6 (Supporting information) shows Pd@TiO2−x(1.5%) exhibits the smallest semicircle of all the electrocatalysts studied with a charge-transfer resistance of 97.83 Ω (Table S1 in Supporting information), implying the fastest charge transfer rate at the electrode and electrolyte interface.

    To further elucidate the impact of differently shaped Pd-loaded oxygen-deficient titanium dioxide on the hydrogen evolution (HER), DFT calculations [66]. Based on the XRD and Raman results, the rutile TiO2(101) surface was employed as a carrier model, as depicted in Fig. S7 (Supporting information), and constructed oxygen vacancies by removing individual oxygen layer of the TiO2(101) surface: 2-coordinated bridging O (O2c) and 3-coordinated in-plane O (O3c), as illustrated in Fig. S8 (Supporting information). The optimized energy clearly indicates that removing 2-coordinated bridging O atoms leads to the formation of more stable oxygen-deficient titanium dioxide (TiO2−x) [67]. Therefore, this structure was applied as the adsorption carrier for Pd, simulating the base configuration of the composite material. The structures of monatomic palladium, cluster palladium and nano-palladium were optimized, the three models are shown in Figs. S9a-c (Supporting information).

    Figs. S9d-f (Supporting information) display differential charge density maps for the three Pd shapes. It is evident that the presence of Pd results in an accumulation of electrons on the carrier, favoring enhanced catalytic performance, consistent with experimental results. Additionally, single-atom Pd induces the highest electron transfer between the single Pd atom and the substrate [68]. Although Pd clusters and Pd nanosheets exhibit larger overall charge transfer to the substrate, the charge transfer to each atom is smaller than that for single-atom Pd. The increased electron transfer in Pd1@TiO2−x activates the reaction, thereby promoting improved HER performance.

    For the HER reaction in acidic electrolyte, the Gibbs free energy of hydrogen adsorption (ΔGH*) is a crucial factor for evaluating hydrogen evolution activity. As shown in Fig. S9g (Supporting information), the ΔGH* for pure TiO2 is −2.77 eV, indicating weak adsorption of H*. TiO2−x exhibits a ΔGH* of −1.16 eV, signifying a substantial reduction in the Gibbs free energy for adsorbing active hydrogen due to the presence of oxygen vacancies. This demonstrates that the presence of oxygen vacancies significantly enhances the interaction between surface oxygen atoms and H*. After introducing metal Pd, a significant reduction in ΔGH* values is observed, indicating that Pd is the primary catalytic active site. Notably, single-atom Pd exhibits the smallest ΔGH* value, at −0.06 eV, lower than Pd clusters (−0.42 eV), Pd nanoparticles (−0.88 eV), and even commercial Pt@C (−0.09 eV). To understand the role of single-atom Pd on the carrier surface, the influence of different Pd shapes on the density of states (PDOS) distribution of d-band electrons on oxygen-deficient titanium dioxide carriers was investigated, as shown in Figs. S9h-j (Supporting information). Pd 3d orbitals, as active sites, influence the formation of the Pd-H bond, a key factor in the HER process [69]. According to the d-band center theory, the coupling of the molecular orbitals of adsorbates with the d orbitals of metal Pd can form two different orbitals: completely occupied bonding orbitals and partially occupied antibonding orbitals. According to the valence bond theory, the occupancy of electrons in antibonding orbitals determines the stability of the system to some extent, that is, the strength of the Pd-H bond. Compared to Pd nanoparticles and Pd clusters, single-atom Pd has a smaller d-band center value, indicating that single-atom Pd exhibits more occupied electronic state [70]. From the figures, it can be observed that single-atom Pd maximally promotes the shift of the d-band center away from the Fermi level, weakens the strength of the Pd-H bond, and lowers the adsorption energy of H. The d-band center of Pd nanosheets is closer to the Fermi level, indicating a stronger interaction with adsorbed H, stronger adsorption energy. This results in the difficulty of H desorption and a larger hydrogen evolution reaction energy barrier [71]. Single-atom Pd has a moderate interaction with H, leading to a d-band center that tends toward zero, making the free energy of adsorption and desorption approach zero, facilitating the progress of the HER.

    This study employed an efficient hydrothermal method to synthesize oxygen-deficient titanium dioxide nanosheets (TiO2−x), and subsequently anchored single-atom Pd on the carrier surface through photo-reduction deposition. Pd@TiO2−x(1.5%) exhibited a low overpotential of 24 mV and a Tafel slope of 41.9 mV/dec, achieving a current density of only 10 mA/cm2 in acidic conditions. Furthermore, the hydrogen evolution reaction (HER) performance remained stable without degradation after continuous operation for 10 h. Density functional theory (DFT) calculations indicated that single-atom Pd possessed a ΔGH* value of −0.06 eV, lower than that of Pd clusters and Pd nanoparticles. The projected density of states (PDOS) results demonstrated that single-atom Pd, by optimizing the electron density around it, forms a strong chemical interaction with the carrier, resulting in a unique electronic structure and more oxygen vacancies. This enhancement promoted the adsorption of protons and the desorption of hydrogen. The presence of oxygen vacancies provided additional active hydrogen adsorption sites, facilitating the generation of active hydrogen and optimizing the material’s activity and conductivity. The synergistic effects of these factors collectively contributed to the improvement of catalytic performance.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Xiaoyu Zhao: Writing – review & editing, Writing – original draft, Supervision, Investigation, Conceptualization. Kai Gao: Methodology, Formal analysis, Data curation. Sen Xue: Visualization, Validation, Data curation. Wei Ran: Software. Rui Liu: Writing – review & editing, Project administration, Funding acquisition.

    This work was financially supported by the Key Project of Tianjin Natural Science Foundation (No. 23JCZDJC00570), Special Funding of China Postdoctoral Science Foundation (No. 2023T160268), China Postdoctoral Science Foundation (No. 2023M741362), the National natural science foundation of China (Nos. 22276209 and 21822608), and Research Center for Eco-Environmental Science (No. RCEES-TDZ-2021-7). R. Liu acknowledges the support from the Youth Innovation Promotion Association of CAS (No. Y2021019).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.110309.


    1. [1]

      Q. Gao, W. Zhang, Z. Shi, et al., Adv. Mater. 31 (2019) 1802880. doi: 10.1002/adma.201802880

    2. [2]

      Z. He, X. Huang, Q. Chen, M. Zhu, J. Colloid Interface Sci. 616 (2022) 279–286. doi: 10.1016/j.jcis.2022.02.073

    3. [3]

      K. Pramoda, C.N.R. Rao, J. Mater, Chem. A 11 (2023) 16933–16962. doi: 10.1039/d3ta01629c

    4. [4]

      M. Muhyuddin, G. Tseberlidis, M. Acciarri, et al., J. Energy Chem. 87 (2023) 256–285. doi: 10.1016/j.jechem.2023.08.011

    5. [5]

      L.N. Shi, L.T. Cui, Y.R. Ji, et al., Coord. Chem. Rev. 469 (2022) 214668. doi: 10.1016/j.ccr.2022.214668

    6. [6]

      X. Zhao, X. Lu, W. Chen, et al., J. Colloid Interface Sci. 659 (2024) 11–20. doi: 10.1117/12.3047938

    7. [7]

      J. Li, Z. Jing, H. Bai, et al., Environ. Chem. Lett. 21 (2023) 2583–2617. doi: 10.1007/s10311-023-01616-z

    8. [8]

      P. Aggarwal, D. Sarkar, K. Awasthi, P.W. Menezes, Coord. Chem. Rev. 452 (2022) 214289. doi: 10.1016/j.ccr.2021.214289

    9. [9]

      C. Lei, W. Zhou, Q. Feng, et al., Nano-Micro Lett. 11 (2019) 1–10. doi: 10.11648/j.sr.20190701.11

    10. [10]

      X. Shi, C. Dai, X. Wang, et al., Nat. Commun. 13 (2022) 1287. doi: 10.1038/s41467-022-28995-1

    11. [11]

      X. Liu, Z. Hao, H. Wang, et al., Proc. Natl. Acad. Sci. U. S. A. 119 (2022) e2119723119. doi: 10.1073/pnas.2119723119

    12. [12]

      H. Zhao, X. Yang, Y. Duan, Z. Shen, ChemCatChem 16 (2024) e202301295. doi: 10.1002/cctc.202301295

    13. [13]

      C.Z. Yuan, K.S. Hui, H. Yin, et al., ACS Mater. Lett. 3 (2021) 752–780. doi: 10.1021/acsmaterialslett.0c00549

    14. [14]

      J. Li, J. Hu, M. Zhang, et al., Nat. Commun. 12 (2021) 3502. doi: 10.1039/d1qo00111f

    15. [15]

      E. Heydari-Bafrooei, S. Askari, Int. J. Hydrog, Energy. 42 (2017) 2961–2969. doi: 10.1016/j.ijhydene.2016.11.101

    16. [16]

      Z. Pu, I.S. Amiinu, R. Cheng, et al., Nano-Micro Lett. 12 (2020) 1–29. doi: 10.1080/14786419.2020.1830398

    17. [17]

      X. Liu, S. Gao, P. Yang, et al., Appl. Mater. Today 13 (2018) 158–165. doi: 10.3390/catal8040158

    18. [18]

      X. Wu, H. Zhang, S. Zuo, et al., Nano-Micro Lett. 13 (2021) 136. doi: 10.1007/s40820-021-00668-6

    19. [19]

      F. Zheng, C. Zhang, X. Gao, et al., Electrochim. Acta 306 (2019) 627–634. doi: 10.1016/j.electacta.2019.03.175

    20. [20]

      N.W.J. Scott, N. Jeddi M.J. Ford, A. Eyles, et al., J. Am. Chem. Soc. 143 (2021) 9682–9693. doi: 10.1021/jacs.1c05294

    21. [21]

      X. Gao, G. Yu, L. Zheng, et al., ACS Appl. Energy Mater. 2 (2019) 966–973. doi: 10.1021/acsaem.8b01783

    22. [22]

      J. Chen, Y. Ha, R. Wang, et al., Nano-Micro Lett. 14 (2022) 186. doi: 10.1007/s40820-022-00933-2

    23. [23]

      Z. Chen, S. Yun, L. Wu, et al., Nano-Micro Lett. 15 (2023) 4. doi: 10.1007/s40820-022-00974-7

    24. [24]

      B. Geng, X. Chu, L. Liu, et al., Chin. Chem. Lett. 35 (2024) 108924. doi: 10.1016/j.cclet.2023.108924

    25. [25]

      H. Yu, Y. Xue, B. Huang, et al., iScience 11 (2019) 31–41. doi: 10.1016/j.isci.2018.12.006

    26. [26]

      J. Zhu, L. An, X. Li, et al., Appl. Surf. Sci. 644 (2024) 158809. doi: 10.1016/j.apsusc.2023.158809

    27. [27]

      B. Guo, Y. Ding, H. Huo, et al., Nano-Micro Lett. 15 (2023) 57. doi: 10.1007/s40820-023-01038-0

    28. [28]

      A.S.B. Mohd-Najib, M. Iqbal, M.B. Zakaria, et al., J. Mater. Chem. A 8 (2020) 19788–19792. doi: 10.1039/d0ta04223d

    29. [29]

      X. Chen, S. Guan, J. Zhou, et al., Catalysts 62 (2023) e202312734.

    30. [30]

      C. Zhai, Y. Chen, X. Huang, et al., Environ. Funct. Mater. 1 (2022) 219–229. doi: 10.1177/0308518x211055181

    31. [31]

      Q. Qiao, Y. Chen, Y. Wang, et al., Chin. Chem. Lett. 34 (2023) 107394. doi: 10.1016/j.cclet.2022.03.117

    32. [32]

      H. Zhang, Q. Liu, Z. Shen, Chin. Chem. Lett. 35 (2024) 108607. doi: 10.1016/j.cclet.2023.108607

    33. [33]

      S.H. Wang, L.N. Chen, P.Y. Zhang, et al., Nano Energy 88 (2021) 106211. doi: 10.1016/j.nanoen.2021.106211

    34. [34]

      X. Zeng, Y. Bai, S.M. Choi, et al., Mater. Today Nano 6 (2019) 100038. doi: 10.1016/j.mtnano.2019.100038

    35. [35]

      C. Li, N.C.S. Selvam, J. Fang, Nano-Micro Lett. 15 (2023) 83. doi: 10.1007/s40820-023-01060-2

    36. [36]

      X. You, D. Zhang, X.G. Zhang, et al., Nano-Micro Lett. 16 (2024) 53. doi: 10.1007/s40820-023-01285-1

    37. [37]

      J. Yang, M. Zhang, M. Chen, et al., Adv. Mater. 35 (2023) 2209885. doi: 10.1002/adma.202209885

    38. [38]

      T. Zhang, T. Yang, G. Qu, et al., J. Energy Chem. 68 (2022) 71–77. doi: 10.1016/j.jechem.2021.10.031

    39. [39]

      T. Zhang, T. Yang, B. Li, S. Wei, W. Gao, J. Alloys Compd. 905 (2022) 164165. doi: 10.1016/j.jallcom.2022.164165

    40. [40]

      Z. Wu, P. Yang, Q. Li, et al., Angew. Chem. Int. Ed. 135 (2023) e202300406. doi: 10.1002/ange.202300406

    41. [41]

      M. Zhou, X. Jiang, W. Kong, et al., Nano-Micro Lett. 15 (2023) 166. doi: 10.1007/s40820-023-01135-0

    42. [42]

      M. Song, Q. Zhang, T. Shen, et al., Chin. Chem. Lett. 35 (2024) 109083. doi: 10.1016/j.cclet.2023.109083

    43. [43]

      A. Zhao, J. Masa, W. Xia, J. Energy Chem. 23 (2014) 701–707. doi: 10.1016/S2095-4956(14)60202-3

    44. [44]

      C. Wang, Q. Zhang, B. Yan, et al., Nano-Micro Lett. 15 (2023) 52. doi: 10.1007/s40820-023-01024-6

    45. [45]

      K. He, Z. Huang, C. Chen, et al., Nano-Micro Lett. 16 (2023) 23.

    46. [46]

      Z. Zeng, S. Küspert, S.E. Balaghi, et al., Small 19 (2023) 2205885. doi: 10.1002/smll.202205885

    47. [47]

      J. Zhang, M. Zhang, W. Li, et al., J. Phys. Chem. C 112 (2008) 19506–19515. doi: 10.1021/jp8036523

    48. [48]

      W. He, X. Zhang, K. Zheng, et al., Angew. Chem. Int. Ed. 62 (2023) e202213365. doi: 10.1002/anie.202213365

    49. [49]

      N. Xu, K. Wang, Y. Zhu, Y. Zhang, Adv. Mater. 35 (2023) e2303173. doi: 10.1002/adma.202303173

    50. [50]

      W. Wu, S. Bu, L. Bai, et al., Nanoscale 15 (2023) 5909–5918. doi: 10.1039/d2nr04952j

    51. [51]

      T. Luo, J. Huang, Y. Hu, et al., Adv. Funct. Mater. 33 (2023) 2213058. doi: 10.1002/adfm.202213058

    52. [52]

      L. Chen, S.I. Allec, M.T. Nguyen, et al., J. Am. Chem. Soc. 145 (2023) 10847–10860. doi: 10.1021/jacs.3c02326

    53. [53]

      C. Wang, X. Liu, J. Wang, et al., Chin. Chem. Lett. 35 (2024) 108739. doi: 10.1016/j.cclet.2023.108739

    54. [54]

      Y. Zhao, L. Tao, Chin. Chem. Lett. 35 (2024) 108571. doi: 10.1016/j.cclet.2023.108571

    55. [55]

      S. Wu, X. Liu, X. Lian, et al., Adv. Mater. 30 (2018) 1802173. doi: 10.1002/adma.201802173

    56. [56]

      S. Wang, T. Ding, T. Liu, et al., Angew. Chem. Int. Ed. 62 (2023) e202218630. doi: 10.1002/anie.202218630

    57. [57]

      E.D. Corro, M. Taravillo, V.G. Baonza, Phys. Rev. B 85 (2012) 033407.

    58. [58]

      S. Wan, Y. Chen, S. Fang, et al., Nat. Mater. 20 (2021) 624–631. doi: 10.1038/s41563-020-00892-2

    59. [59]

      K.M. Naik, E. Higuchi, H. Inoue, Nanoscale 12 (2020) 11055–11062. doi: 10.1039/d0nr02092c

    60. [60]

      M. Zhang, W. Guo, Y. Chen, et al., Chin. Chem. Lett. 34 (2023) 108229. doi: 10.1016/j.cclet.2023.108229

    61. [61]

      H. Wei, X. Li, B. Deng, et al., Chin. J. Catal. 43 (2022) 1058–1065. doi: 10.1016/S1872-2067(21)63968-2

    62. [62]

      J.Y. Yue, X.L. Ding, L.P. Song, et al., Micropor. Mesopor. Mat. 5 (2021) 3859–3866. doi: 10.1039/d1qm00314c

    63. [63]

      F. Meng, X. Qin, L. Yang, et al., Small 18 (2022) 2203283. doi: 10.1002/smll.202203283

    64. [64]

      Z. Wei, H. Wang, C. Zhang, et al., Angew. Chem. Int. Ed. 60 (2021) 16622–16627. doi: 10.1002/anie.202104856

    65. [65]

      Y. Chen, L. Soler, M. Armengol-Profitós, et al., Appl. Catal. B 309 (2022) 121275. doi: 10.1016/j.apcatb.2022.121275

    66. [66]

      W. He, X. Zhang, K. Zheng, et al., Angew. Chem. Int. Ed. 62 (2023) e202213365. doi: 10.1002/anie.202213365

    67. [67]

      J. Cui, Z. Hao, Y. Wang, et al., Chem. Eng. J. 446 (2022) 136989. doi: 10.1016/j.cej.2022.136989

    68. [68]

      O.A. Qamar, F. Jamil, M. Hussain, et al., Chem. Eng. J. 460 (2023) 141734. doi: 10.1016/j.cej.2023.141734

    69. [69]

      X. Chen, M. Peng, X. Cai, et al., Nat. Commun. 12 (2021) 2664. doi: 10.1158/1078-0432.ccr-21-0936

    70. [70]

      S. Ji, Y. Chen, G. Zhao, et al., Appl. Catal. B 304 (2022) 120922. doi: 10.1016/j.apcatb.2021.120922

    71. [71]

      S. Weon, M. J Suh, C. Chu, et al., Environ. Sci. Technol. 1 (2021) 512–522. doi: 10.1021/acsestengg.0c00210

  • Figure 1  (a) XRD patterns of TiO2-x, Pd@TiO2-x(1.5%), Pd@TiO2-x(3%), Pd@TiO2-x(6%), Pd@TiO2-x(10%). (b) TEM images of TiO2−x. (c, d) TEM images of different crystal forms of TiO2−x. (e-h) TEM images of Pd@TiO2−x(1.5%), Pd@TiO2−x(3%), Pd@TiO2−x(6%), Pd@TiO2−x(10%). (i) HAADF-STEM image of Pd@TiO2−x(1.5%).

    Figure 2  (a, b) High-resolution XPS Ti 2p and O 1s spectrum of TiO2−x. (c) Raman spectrums of TiO2−x, Pd@TiO2−x(1.5%), Pd@TiO2−x(3%), Pd@TiO2−x(6%), Pd@TiO2−x(10%). (d) EPR spectra of different catalyst.

    Figure 3  (a, b) HER polarization curves and overpotential histogram of as-prepared samples in 50 mmol/L H2SO4, respectively. (c) Schematic illustration of the HER mechanism on Pd@TiO2−x(1.5%). (d, e) Tafel plots and Slope histogram of prepared samples in 50 mmol/L H2SO4, respectively. (f) TOF curves with respect to Pd or Pt atoms and mass activities (based on Pd or Pt mass loading) at η = 50 mV of catalysts. (g) Corresponding current density scan rate curves to estimate the Cdl, respectively. (h) Cycling stability tests of Pd@TiO2−x(1.5%) by 1000 cycles. (i) Long-term chronoamperometry measurements at a current density of 10 mA/cm2 of Pd@TiO2−x(1.5%).

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  161
  • HTML全文浏览量:  3
文章相关
  • 发布日期:  2025-06-15
  • 收稿日期:  2024-02-19
  • 接受日期:  2024-07-30
  • 修回日期:  2024-06-25
  • 网络出版日期:  2024-07-31
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章