Next Article in Journal
Fabrication of CoP@P, N-CNTs-Deposited Nickel Foam for Energy-Efficient Hydrogen Generation via Electrocatalytic Urea Oxidation
Previous Article in Journal
Chitosan-Modified SBA-15 as a Support for Transition Metal Catalysts in Cyclohexane Oxidation and Photocatalytic Hydrogen Evolution
Previous Article in Special Issue
Active Ag-, Fe-, and AC-Modified TiO2 Mesoporous Photocatalysts for Anionic and Cationic Dye Degradation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Influence of Defect Engineering on the Electronic Structure of Active Centers on the Catalyst Surface

1
School of Chemistry, Beihang University, Beijing 100191, China
2
International Institute for Interdisciplinary and Frontiers, Beihang University, Beijing 100191, China
3
Research Institute of Aero-Engine, Beihang University, Beijing 100191, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(7), 651; https://doi.org/10.3390/catal15070651
Submission received: 30 May 2025 / Revised: 28 June 2025 / Accepted: 30 June 2025 / Published: 3 July 2025

Abstract

Defect engineering has recently emerged as a cutting-edge discipline for precise modulation of electronic structures in nanomaterials, shifting the paradigm in nanoscience from passive ‘inherent defect tolerance’ to proactive ‘defect-controlled design’. The deliberate introduction of defect—including vacancies, dopants, and interfaces—breaks the rigid symmetry of crystalline lattices, enabling new pathways for optimizing catalysis performance. This review systematically summarizes the mechanisms underlying defect-mediated electronic structure at active sites regulation, including (1) reconstruction of the electronic density of states, (2) tuning of coordination microenvironments, (3) charge transfer and localization effects, (4) spin-state and magnetic coupling modulation, and (5) dynamic defect and interface engineering. These mechanisms elucidate how defect-induced electronic restructuring governs catalytic activity and selectivity. We further assess advanced characterization techniques and computational methodologies for probing defects-induced electronic states, offering deeper mechanistic insights at atomic scales. Finally, we highlight recent breakthroughs in defect-engineered nanomaterials for catalytic applications, including hydrogen evolution reaction (HER), oxygen evolution reaction (OER) and beyond, while discussing existing challenges in scalability, defect stability, and structure–property causality. This review aims to provide actionable principles for the rational design of defects to tailor electronic structures toward next-generation energy technologies.

Graphical Abstract

1. Introduction

With the rapid advancement of nanotechnology and characterization techniques, materials science is transitioning into an era of atomic-scale design. The “perfection” of traditional crystalline materials is being reexamined—the deliberate introduction of specific defects not only preserves material functionality but also unlocks novel electronic states [1]. Defect engineering, the methodology for purposefully introducing defects such as vacancies, dopant, and grain boundaries, is redefining the foundational paradigms in catalysis, optoelectronic devices, and energy storage [2,3,4,5,6]. As a core strategy for electronic structure modulation, its essence lies in leveraging defect-induced local atomic rearrangements and electronic state reconstructions to overcome the limitation of ideal crystalline lattices, thereby granting multidimensional degrees of freedom for the design of novel material systems [7,8,9]. This review systematically explores the mechanisms of defect-mediated electronic structure regulation, synthesizes recent insights into the defect-electronic structure–property triad, and establishes a conceptual framework for the rational design of defect at specific electronic architectures.
In recent years, defect-mediated electronic structure modulation has been increasingly recognized as a multiscale and synergistic phenomenon. At the atomic scale, vacancies or dopant atoms can directly influence the electron donor/acceptor characteristics of active sites by altering localized charge density and bonding strength. For example, Fe doping combined with Co vacancies in CoSe2 nanobelts regulates the electronic structure of Co2 sites, reconstructing the electronic density of states near the Fermi level [10]. At the mesoscopic scale, extended defects such as interfaces and dislocations influence macroscopic carrier transport pathways via mechanisms including space-charge regions and quantum confinement effects. An example is the WS2–WO3 heterostructure, where abundant heterogeneous interfaces generate a strong interfacial electric field. These fields elevate the d-band center of W atoms and enhance the adsorption of intermediates [11]. Dynamically, the evolution of defect structures during operation triggers real-time electronic state reconstruction, which critically determines catalyst activity and stability. In the case of self-supported coral-like Ce2W2O9 arrays, the bond structure of Ce atoms was regulated by the W defects formed through in situ electrochemical reconstruction, and the adsorption of intermediates by Ce active sites was optimized [12].
This review elucidates the physicochemical mechanisms by which defect engineering modulates electronic structures, organized through a tripartite analytical framework:
  • Multidimensional regulatory strategies: We systematically dissect how defect engineering modulate the electronic structures of active sites, including electronic density-of-states reconstruction, coordination microenvironment tuning, charge transfer dynamics, and localization effects. Based on this, we summarized the main defect types in nanocatalysts and their regulatory effects on the electronic structure, as shown in Table 1.
  • Multiscale characterization techniques: We critically examined advanced techniques used for probing defect-electronic structure–property relationships. These includes defect-specific spectroscopies, electronic structure analyses and computational modeling approaches.
  • Performance benchmarks: We highlight representative catalytic systems where defect-mediated electronic tailoring leads to enhanced performance, with a focused on hydrogen evolution reaction (HER), oxygen evolution reaction (OER) and other relevant reaction systems.
This review focuses on the regulation mechanism, characterization and effect of defect engineering strategies on the electronic structure on the surface of nanocatalysts. It provides a detailed analysis of how defect engineering regulates the electronic structure and comprehensively summarizes research strategies such as defect characterization, electronic structure characterization, theoretical calculation and machine learning. This review also evaluates the significant contributions of the defect engineering regulation electronic structure strategy in fields such as HER and OER. By integrating regulation strategies, characterization techniques, computational modeling and experimental outcomes, this paper establishes the general design principles for defect-driven electronic modulation, aiming to provide a unified perspective for the rational design of the next generation of functional materials.

2. Mechanisms Underlying Defect Engineering-Mediated Modulation of Electronic Structures at Active Sites

2.1. Electronic Density of States Reconstruction

2.1.1. Defect Energy Level Engineering

Defect energy level engineering enables precise modulation of electronic structure through mechanisms including bandgap tuning, carrier dynamics regulation, and interfacial electron transfer optimization. By tailoring band structures to enhance light absorption and charge separation kinetics, coupled with the modification of local electronic environments to generate highly active catalytic sites, this approach offers innovative strategies for designing high-performance photocatalysts. The deliberate introduction of defect energy levels has been successfully applied in a range of applications including photocatalysis, molecular oxygen activation, and pollutant degradation, firmly establishing defect engineering as a key methodology for achieving atomic-level control over electronic architectures.
The strategic tuning of oxygen vacancy (OV) concentrations has emerged as a critical factor in optimizing charge carrier dynamics and separation efficiency in S-scheme heterojunction photocatalysts. A notable example was reported by Xu et al., who constructed a WO3−x/In2S3 heterostructure and elucidated the critical role of defect energy levels in prolonging carrier lifetimes [13]. Moderate OV concentrations introduced intermediate defect states within the WO3−x bandgap, which acted as electron reservoirs. These states significantly extended carrier lifetimes by facilitating the rapid electron trapping and delayed release of photogenerated electrons. X-ray photoelectron spectroscopy (XPS) analysis revealed negative shifts in W 4f and O 1s binding energies, along with changes in the W5+/W6+ ratio with increasing calcination temperature, confirming OV-induced redistribution of electronic density. Density functional theory (DFT) calculations further demonstrated that progressive OV incorporation led to the formation of shallow-to-deep defect levels, changing the bandgap of WO3. Density of states (DOS) analysis corroborated this effect, showing an increasing depth of defect levels with higher OV concentrations (Figure 1). Femtosecond transient absorption spectroscopy (fs-TAS) provided direct kinetic evidence that optimal OV concentrations prolonged carrier lifetimes by functioning as electron traps, whereas excessive deep defect states (e.g., in WO3−x-500) that quenched the deeply trapped electrons and shortened lifetimes. In situ irradiation XPS (ISI-XPS) and fs-TAS measurements collectively confirmed that intermediate defect levels—introduced by controlled OVs—play a critical role in electron storage and delayed release, thereby enhancing charge carrier separation. Notably, the optimized defective S-scheme heterojunction exhibited exceptional CO2 photoreduction performance, achieving nearly 100% CO selectivity. These findings underscore the efficacy of defect-level engineering in tailoring charge transfer kinetics and photocatalytic efficiency.
As another example of the energy level regulation electronic structure induced by oxygen vacancies, Yang et al. engineered a Ni2P/Bi3O4Br-OVs heterojunction system, in which OVs introduced localized defect energy levels within the Bi3O4Br bandgap, substantially optimizing its electronic properties [14]. The introduced OVs generated a deep defect level associated with Bi 6p orbitals of Bi(3−x)+ near the mid-gap region, facilitating excitation under low-energy photons and serving as electron-trapping centers that improved O2 activation. Electron paramagnetic resonance (EPR) and X-ray photoelectron spectroscopy (XPS) analyses confirmed the presence of OVs and their modulatory effects on Bi3+ electronic states, as indicated by negative binding energy shifts in Bi 4f and Br 3d binding energies, indicating surface charge redistribution. Theoretical simulations further revealed that dual active sites (OVs and Ni2P) synergistically enhanced oxygen activation and charge separation, generating holes (h+) while converting O2 into superoxide radicals (·O2), which serve as pivotal reactive intermediates in the photocatalytic oxidation process.
Introducing sulfur vacancies (SVs) into two-dimensional nanostructures has emerged as an effective strategy to optimize piezocatalytic performance via band structure engineering and charge carrier regulation. Ma et al. synthesized ultrathin 2D sulfur vacancy-engineered (SV-engineered) Cu@SnS2−x nanosheets through a heterovalent substitution strategy, wherein Sn4+ sites were partially replaced by Cu2+ [15]. The incorporation of Cu2+ preferentially induced controlled SVs within the SnS2 lattice, which significantly modified its electronic structure and charge transfer kinetics. These SVs narrowed the bandgap of SnS2 from 2.16 eV (pristine) to 1.62 eV (Cu@SnS2−x), thereby facilitating electron-hole (e-h+) separation and enhancing sonocatalytic efficiency. Combined experimental and theoretical analyses elucidated the role of SVs in electronic modulation. X-ray photoelectron spectroscopy (XPS) demonstrated negative shifts in binding energies for both S and Sn, indicative of localized electron density redistribution around Sn atoms. DFT calculations revealed an upward shift in the Fermi level of SnS2 following Cu doping, suggesting the formation of defect states t the lower region of CB edge. These states promoted electron transfer, improving the separation of ultrasonically generated carriers (Figure 2a,b). Piezoresponse force microscopy (PFM) further confirmed the defect-mediated charge separation mechanism, as evidenced by a pronounced increase in the piezoelectric coefficient signal amplitude for Cu@SnS2−x. Notably, the Cu@SnS2−x nanosheets exhibited ultrasound-enhanced, tumor microenvironment (TME)-responsive Fenton-like catalytic activity and glutathione depletion capability, enabling efficient tumor cell ablation via biocompatible piezocatalytic therapy. This study highlights the biomedical promise of defect engineering in designing electronic structures for targeted therapeutic applications.
For metal-free nanomaterials, the introduction of midgap states through heteroatom doping can effectively inhibit charge recombination and improve the migration of photogenerated carriers. Among various doping strategies, nitrogen doping has been particularly effective in inducing defect energy levels, thereby facilitating the spatial separation of photogenerated electron-hole pairs. Xing et al. reported the synthesis of nitrogen-doped porous thin-walled graphitic carbon nitride (g-C3N4) nanotubes (denoted as CNx) [16]. The incorporation of nitrogen atoms introduced midgap energy levels within the electronic band structure, resulting in a reduction in the bandgap from 2.46 eV to 2.28 eV for the CN0.75 composition. These nitrogen-induced midgap states enhance visible light harvesting by enabling the absorption of lower-energy photons and act as effective charge separation centers rather than recombination centers. Consequently, CN0.75 exhibited remarkable photocatalytic performance for tetracycline degradation, achieving a degradation efficiency of 93.0% within 60 min. Moreover, the material demonstrated excellent stability, with only a 4.1% decrease in photocatalytic efficiency after five consecutive cycles. This study underscores the significance of defect-level engineering as a general and effective strategy for optimizing the performance of photocatalysts in applications related to energy conversion and environmental remediation.

2.1.2. Fermi Level Shift

Modulating the Fermi level represents one of the central mechanisms by which defect engineering regulates electronic structures. By introducing defect configurations such as dopants and vacancies, the Fermi level undergoes precise shifts, which in turn alters the electronic interactions between reactants and the catalyst surface. This approach enables systematic tuning of adsorption energetics and charge transfer dynamics at active sites, offering a strategic pathway to enhance catalytic selectivity and activity for targeted reactions.
Interface engineering has become a powerful strategy for tuning Fermi levels and promoting directional charge transfer across catalytic heterojunctions. In this context, Lin et al. constructed a multi-armed MoSe2/CdS S-scheme heterojunction featuring strong Mo–S coupling and tunable selenium vacancy (Vse) and Mo5+ concentrations [17]. The synergistic interplay between Vse vacancies and Mo5+ ions served as a key mechanism for Fermi level adjustment. By modulating the synthesis temperature (160–220 °C), the concentration of donor and acceptor in MoSe2 dynamically evolved. As temperature increased, Vse concentration decreased, causing a downward shift in MoSe2’s EF. However, the simultaneous reduction in Mo5+ concentration counteracted this trend, resulting in a nonlinear EF response (Figure 3): EF initially decreased (ΔEF = −0.08 eV from 160 °C to 180 °C) and then increased (ΔEF = +0.05 eV from 180 °C to 220 °C). This nonlinear relationship modulated the EF offset between CdS and MoSe2, creating a robust built-in electric field. The strengthened field drove spatial separation of high-reducibility photoelectrons in CdS’s conduction band (CB) and high-oxidability photo-holes in MoSe2’s valence band (VB), achieving an outstanding hydrogen evolution activity of 52.62 mmol g−1 h−1 under visible light, corresponding to an apparent quantum efficiency of 34.8% at 400 nm. This work establishes defect-concentration-driven EF modulation as a universal strategy for designing high-performance heterojunction photocatalysts.
The strategic engineering of interfacial kinetics through oxygen vacancies (OVs) has emerged as a powerful approach to control charge carrier dynamics in photocatalytic heterostructures. As demonstrated by Liu et al., the synthesis of a BiVO4/BiOBr-Ov (BVB-Ov) heterostructure revealed that Ov introduction elevates the Fermi level of BiOBr-Ov above that of BiVO4, thereby reconfiguring the interfacial electron transfer pathway [18]. Notably, the elevated Fermi level of BiOBr-Ov induced a directional migration of photogenerated electrons from BiOBr-Ov to BiVO4 upon heterojunction formation. This process triggered downward band bending at the BiVO4 surface and established a pronounced built-in electric field. The resultant band alignment optimization facilitated spatially directed charge transfer across the interface, leading to a remarkable enhancement in carrier migration efficiency. Crucially, the presence of oxygen vacancies mediated a transition from Type-II to Z-scheme charge transfer mechanism, subsequently promoting the generation of oxidative •O2 radicals. The composite containing 20% BVB-Ov exhibited optimal photocatalytic activity, achieving 91% degradation of oxytetracycline (OTC) within 60 min under visible light irradiation (Figure 4). This study highlights Fermi level modulation via oxygen vacancies as an effective strategy for heterojunction-type conversion, providing a viable route to enhance photocatalytic performance through controlled interfacial energetics.

2.2. Coordination Environment Engineering

2.2.1. Low-Coordination Environments Induce the Exposure of Catalytically Active Sites

In low-coordination defect engineering, catalytic activity can be enhanced by exposing highly active sites and reconstructing localized electronic states to optimize adsorption energetics of reaction intermediates. Yu et al. developed amorphous PdSe2 nanoparticles (a-PdSe2 NPs) with disordered atomic arrangements, which generated low-coordinated Pd sites (Figure 5), that break the symmetric coordination constraints inherent to crystalline PdSe2 [19]. DFT calculations reveal that the Pd-d and Se-p orbital states in a-PdSe2 shift closer to the Fermi level compared to crystalline PdSe2, resulting in the formation of continuous conduction bands, metallic behavior, and pronounced electronic state occupancy near the Fermi level. This electronic reconfiguration optimizes the formation energy of the *OOH intermediate, effectively suppressing O–O bond cleavage and promoting selective H2O2 generation. Experimentally, a-PdSe2/C maintains >90% H2O2 selectivity across a broad potential window (0.3–0.6 V vs. reversible hydrogen electrode, RHE) in 0.1 M KOH, surpassing the performance of crystalline Pd-based catalysts. X-ray photoelectron spectroscopy (XPS) and X-ray absorption fine structure (XAFS) analyses further confirmed localized charge redistribution at undercoordinated Pd sites. This work establishes a novel paradigm for tailoring electronic structures via amorphous defect engineering and elucidates the structure–performance relationship between dynamic coordination defects and catalytic activity. These findings provide critical insights for the precision design of catalysts with highly exposed, electronically optimized active sites.
The synergy between amorphization and oxygen doping presents a powerful strategy for modulating low-coordination environments and optimizing electronic states in electrocatalysts. Zhao et al. systematically demonstrated this strategy by transforming crystalline PtTe2 nanosheets (c-PtTe2 NSs) into oxygen-doped amorphous structures (a-PtTe2 NSs) via an oxygen-assisted calcination method [20]. During the crystalline-to-amorphous transition, oxygen doping disrupts the long-range order of c-PtTe2 NSs, inducing structural amorphization and exposing reconfigured interfacial Pt atoms. Partial density of states (PDOS) analysis reveals that the d-band center of a-PtTe2 NSs shifts upward relative to c-PtTe2 NSs, strengthening hydrogen intermediate (H*) adsorption and facilitating rapid proton transfer kinetics. Under an applied electric field, electron accumulation at the exposed Pt sites and their expanded orbital volume further enhanced H* conversion efficiency. Consequently, a-PtTe2 NSs exhibit optimized Gibbs free energy for H* adsorption and exceptional HER performance, including an ultralow initiation potential (~0 mV vs. RHE), an overpotential of 14 mV at 10 mA cm−2, and a Tafel slope of 24.5 mV dec−1. This study demonstrates that controlled amorphization of 2D layered materials circumvents the inherent crystallographic constraints on active site exposure, offering a novel strategy to design catalysts with high-density, undercoordinated active centers.
Tailoring coordination environments through defect engineering stabilizes unconventional active site geometries, thereby enhancing catalytic performance in complex reactions such as nitrogen reduction. Wang et al. engineered a defect-rich photocatalyst composed of Fe single atoms anchored in hollow BiOBr microtubes, where synergistic oxygen (VO) and bromine vacancies (VBr) generate low-coordinated FeO5 active centers, significantly boosting nitrogen reduction reaction (NRR) efficiency [21]. In the pristine FeSA-Bi/BiOBr-VBr system, each Fe atom is coordinated by four oxygen atoms from [Bi2O2]2+ layers (Figure 6b). Introducing dual vacancies disrupts this symmetry, resulting in a pentacoordinated Fe center bound to four oxygen atoms from [Bi2O2]2+ and one oxygen atom located in the Br vacancy (denoted as FeSA-Bi/h-BiOBr-VO,Br; Figure 6a). This low-coordination environment induces localized polar charge redistribution on Fe sites, accelerating electron transfer from Fe to N2 and lowering the activation energy of the potential-determining step (N2 → NH3). Consequently, the optimized catalyst achieves enhanced NH3 production with high intrinsic activity and selectivity. This work provides a universal framework for designing defect-driven catalysts via precise coordination environment modulation.
The integration of undercoordinated multimetallic sites within well-defined frameworks offers a promising strategy for steering complex reaction pathways such as CO2 photoreduction. Wang et al. engineered a multi-atomic catalyst (MAC) with undercoordinated triangular Cu3 sites (denoted as Cu3 MAC) via a “pre-locking and nanoconfinement polymerization” strategy, enabling highly selective CO2 photoreduction to ethanol [22]. EXAFS analysis confirmed the coordination environment: the Cu–N and Cu–Cu coordination numbers for triangular Cu3 MAC were determined as 1.95 and 1.83, respectively, indicating the formation of a low-coordination triangular Cu3 geometry. Under light irradiation, these Cu3 motifs facilitate electron accumulation, generating a high proportion of Cu+ active species critical for asymmetric C–C coupling during CO2 reduction. The Cu3 MAC exhibits exceptional photocatalytic performance for CO2-to-ethanol conversion, attributed to the synergistic interplay of high-density Cu3 sites with mixed Cu+/Cu2+ oxidation states, which enhance CO2 chemisorption, accelerate charge separation, and stabilize the *CHCHOH intermediate. This “geometric-electronic dual modulation” strategy establishes a versatile framework for designing selective multicarbon-product catalysts.
Inducing low-coordination sites by constructing lattice mismatch interfaces represents an effective strategy for optimizing the electronic structure and improving catalytic activity. Zhang et al. engineered a self-supporting NiO/Co3O4 hybrid catalyst with abundant heterointerfaces, where low-coordination atoms (LCAs) at lattice-mismatched interfaces optimize electronic configurations and expose active sites for the OER [23]. The LCAs mediate electronic modulation, shifting Co and Ni cations to the favorable oxidation species in the intermediate spin states. DFT calculations reveal that interfacial LCAs downshift the Co d-band center below the Fermi level, thereby weakening oxygen intermediate adsorption strength and enhancing intrinsic activity. The hybrid catalyst achieves an OER overpotential of 262 mV at 10 mA cm−2, outperforming NiO (456 mV), Co3O4 (490 mV), Ni foam (680 mV), and commercial IrO2 (361 mV), with a Tafel slope of 58 mV dec−1. This study exemplifies interface defect engineering: precise control over LCA spatial distribution and electronic density of states circumvents adsorption–desorption limitations in conventional metal oxides, paving the way for efficient, stable OER electrocatalysts.

2.2.2. Lattice Distortion Engineering for Electronic Structure Modulation

Lattice distortion has emerged as a vital tool for inducing local strain fields and electronic configurations at catalytic sites. By altering bond lengths, and reducing coordination numbers of active sites, strain engineering effectively elevates d-band centers, and modulates the electronic density of states, thereby enhancing catalytic performance.
In a representative study, Liu et al. engineered a strained RuO2−x system via molten salt-assisted quenching, generating localized tensile strain on RuO2−x surfaces, and incorporating Sr/Ta dopants to optimize its electronic structure [24]. The tensile strain induces spatial elongation of Ru–O bonds and reduced covalency, thereby suppressing lattice oxygen participation and mitigating structural degradation (Figure 7). Concurrently, Sr–Ta co-doping synergistically modulates the electronic environment, optimizing deprotonation kinetics at oxygen sites and tuning intermediate adsorption energetics at Ru sites, which collectively lower reaction energy barriers. The strained catalyst achieves an oxygen evolution reaction (OER) overpotential of 166 mV at 10 mA cm−2 in 0.5 M H2SO4, with stability metrics exceeding pristine RuO2 by an order of magnitude. Long-term stability evaluations in single-cell and proton exchange membrane (PEM) electrolyzers further validate its exceptional durability and activity, surpassing most state-of-the-art acidic OER catalysts.
Strain engineering at the nanoscale offers a powerful strategy for reconfiguring surface atomic structures and tuning electronic states, particularly in transition metals with intrinsically flexible bonding networks. Li et al. synthesized lattice-distorted Cu containing abundant dislocation-nanotwinned structures (DNTs) via electrochemical reduction of Cu2O, generating strained nanostructures enriched with dislocations, axial-shear strain, and reduced surface coordination numbers (CN) [25]. The elecated tensile strain and decreased CN reduce Cu–Cu electron sharing, shifting the d-band center upward and lowering the hydrogen adsorption ΔGH*. At a coordination number of 5, ΔGH* on Cu(111)/(110) surfaces approaches zero, indicating optimal hydrogen evolution reaction (HER) activity. Although dislocation-twinned Cu initially matches Pt/C in performance at low overpotentials due to limited active site accessibility, it surpasses Pt/C under high-current-density conditions, delivering an overpotential of 301 mV (vs. 425 mV for Pt/C) at 500 mA cm−2. This strain engineering strategy establishes Cu as a scalable, cost-effective HER catalyst for industrial-scale applications.
Carbon-based nanomaterials with engineered strain and defect sites offer versatile platforms for modulating reactivity, especially in oxidation and halogenation reactions. Zhang et al. synthesized defect-rich, strain-engineered carbon nanotubes (CNTs) characterized by enlarged diameters, enhanced porosity, and expanded interlayer spacing, which destabilize sp2-hybridized C–C bonds within the CNT framework [26]. This structural modification introduces localized tensile strain and modulates sp2-hybridized C–C bonding, thereby promoting electron transitions from 1s to π* orbitals and enhancing surface reactivity. The elongation of sp2 bonds at strain sites enhances electronic interactions with chlorine molecules via an initial Yeager-type adsorption mechanism. As a result, this strain-driven process elevates ClO• selectivity from 38.8% (conventional defective CNTs) to 87.5%, enabling deep mineralization of 2,4-dichlorophenol (2,4-DCP). The study underscores the pivotal role of strain fields in governing adsorption configurations and regulating chlorination pathways.
Tuning lattice distortion and cation site occupation in mixed-metal oxides can significantly influence defect formation and catalytic functionality. Zhang et al. synthesized an inverse spinel MnOx/Fe1.8O2.4−x via a sol–gel method, wherein Mn3+ preferentially occupies tetrahedral sites, inducing lattice distortion and facilitating the generation of highly active oxygen vacancy (OV) sites near Fe centers [27]. These OVs enhance the adsorption capacity of Fe3+ for oxygen atoms, promoting the adsorption and activation of oxygen-containing molecules. The resulting catalyst demonstrates exceptional selectivity for hydrogenating unsaturated aldehydes to their corresponding unsaturated alcohols. For α,β-unsaturated fatty aldehydes, butenal is converted to butenol with a selectivity of 99.9%. Similarly, the hydrogenation of α,β-unsaturated aromatic aldehyde cinnamaldehyde achieves 98.6% selectivity. In bio-based furfural hydrogenation, the catalyst attains a conversion rate of 93.5% with 97% selectivity after 2 h. At 200 °C, the hydrogenation of the unsaturated carbonyl compound trans-benzyl acetone also achieves 99.9% selectivity.

2.3. Charge Transfer and Localization Effects

2.3.1. Defects as Charge Traps

Defect-induced charge trapping is a critical mechanism for achieving spatial charge separation and prolonging carrier lifetimes in photocatalytic systems. Engineered defects can drive spatial charge separation, and facilitate directional migration of photogenerated carriers by acting as electron or hole traps. By suppressing electron-hole recombination and prolonging charge carrier lifetimes, these defects further generate localized energy levels that alter the band structure, thereby significantly enhancing catalytic activity.
In a representative study, Ma et al. synthesized cyano-defect-enriched graphitic carbon nitride (g-C3N4) with high surface area (denoted as NCN) and coupled it with metallic 1T-WS2 cocatalysts to construct a 1T-WS2/NCN photocatalytic system [28]. The strategically introduced cyano defects in NCN act as electron-trapping centers, promoting spatial segregation of photogenerated charge carriers. This defect-mediated charge separation facilitates the directional migration of carriers, enabling efficient electron excitation from the valence band to unoccupied conduction band orbitals and generating unpaired electrons. Upon photoexcitation, the strongly bound excitons formed within 1T-WS2 layers interact with these trapped electrons to form trions, which localize at catalytically active edge sites. This charge concentration mechanism culminates in the efficient reduction of H+ to H2. The optimized 1T-WS2/NCN system exhibits a remarkable hydrogen evolution rate of 5769.3 μmol·g−1 under 180 min of visible-light irradiation, representing a 90.6-fold enhancement compared to pristine g-C3N4.
Single-atom catalysts (SACs) are profoundly enhanced by defect engineering strategies that create localized trap states, enabling precise regulation of carrier dynamics and modulation of the electronic band structure. Zhang et al. engineered a single indium-atom catalyst (In-Nv-CNF) by introducing N2C vacancies into carbon nitride (CN) frameworks [29]. These vacancies induce intraband defect states that function as electron traps, effectively suppressing electron-hole recombination and extending hole lifetimes (Figure 8a). As a result, the In-Nv-CNF achieves a photogenerated hole concentration of 1.12 × 1015 spins·mm−3, surpassing pristine CN and indium nanoparticle-decorated CN (In NPs/CN) by 74.7- and 24.9-fold, respectively. The optimized In-Nv-CNF demonstrates a 50-fold enhancement in photo-oxidative degradation rates compared to pristine CN, achieving efficient decomposition of tetracycline and ciprofloxacin under visible-light irradiation.
Asymmetric active sites, coupled with defect-induced charge localization, can synergistically enhance multielectron catalytic pathways such as C–C coupling in CO2 reduction. Su et al. engineered asymmetric Cu-Cuδ+-Wδ+ triatomic sites on oxygen-vacancy-rich CuWO4 (CWO-OVs) for photodriven CO2 C–C coupling [30]. OVs functions as electron traps to prolong charge carrier lifetimes and enable a continuous proton-coupled electron transfer (PCET) process (Figure 8b). These synergistic effects drove propionic acid production at 86.46 μmol g−1 h−1 with 89.27% electron selectivity, surpassing pristine CWO. This work demonstrates how defect-mediated charge localization and site asymmetry synergize to optimize multicarbon synthesis.
Embedding catalytic components within defect-rich carbon matrices can significantly enhance interfacial charge transfer and reactive species generation through synergistic electronic modulation. Bai et al. synthesized Co-substituted Bi3O4Cl quantum dots (CBO) embedded in defect-rich biochar (NBC) to create NCBO. Carbon defects in NBC act as electron traps, forming localized energy levels that redistribute electron density and modify the band structure [31]. These defect-mediated changes with quantum confinement enhance interfacial charge transfer and regenerate Co/Bi active sites, enabling efficient •OH and 1O2 generation for pollutant mineralization. NCBO achieves complete detoxification of TeCG via a self-sustaining redox cycle, offering a novel strategy for advanced oxidation processes.
Cations and anions, functioning as positive and negative charge traps, respectively, can significantly enhance charge separation efficiency and catalytic performance. Zhang et al. engineered Cu-induced adaptive sulfur vacancies on ZnIn2S4 nanosheets to regulate charge separation kinetics and construct hydrogen-migration gradient channels [32]. In this system, Cu dopants act as hole traps, while S vacancies serve as electron traps, synergistically suppressing electron-hole recombination. The distinct sulfur atom sites establish a ΔGH* gradient, enabling surface-confined hydrogen migration driven by photothermal effects without interlayer transport. The optimized catalyst, with 5 mol% Cu confined in ZnIn2S4, achieves a hydrogen evolution reaction (HER) rate of 9.8647 mmol g−1 h−1 (14.8-fold enhancement over pristine ZnIn2S4) and an apparent quantum efficiency (AQE) of 37.11% at 420 nm, highlighting the pivotal role of defect engineering in tailoring reaction dynamics and catalytic activity.

2.3.2. The Metal-Support Interaction

The metal-support interaction (EMSI) represents a crucial mechanism that can regulate the electronic structure of metal active sites, inducing electron redistribution between the metal sites and support atoms. This interaction effectively enhances the reactivity of the metal sites and modulates the adsorption energy of intermediate species, thereby tuning their adsorption, improving catalytic activity, and controlling reaction pathways.
Xie et al. utilized this principle to synthesize Mo-doped VO2 nanobelts, where atomically dispersed Mo was embedded within the VO2 matrix to construct a high-efficiency electrocatalyst for NRR [33]. The system utilizes VO2 as a support to strongly adsorb H-adsorbed N2, providing an N2 source to suppress the competing HER. Critically, electronic metal-support interaction between the support and Mo facilitates electron transfer from Mo to the support, generating electron-deficient regions around Mo. These regions accommodate the lone electron pairs of N2 molecules, effectively activating the N≡N bond and reducing the energy barrier for the initial hydrogenation step. The electron-deficient sites also inhibit H+ adsorption, thereby enhancing the Faradaic efficiency (FE) of NRR while suppressing HER. The Mo/VO2 catalyst achieves a remarkable NH3 yield of 190.1 μg NH3 mg cat.−1 h−1 and an FE of 32.4% at −0.5 V vs. RHE, representing 10.8-fold and 2.8-fold improvements over pristine VO2, respectively. This work provides a strategic approach for designing high-efficiency electrocatalytic NRR catalysts.
Plasma-assisted defect engineering can significantly enhance metal-support interactions by introducing surface vacancies and heteroatom dopants, which jointly modulate interfacial charge transfer and catalytic stability. Zhu et al. developed Ir@Sr-p-TiO2 nanowires via plasma treatment, integrating Ir nanoparticles and Sr single atoms for bifunctional OER and HER [34]. The enhanced activity stems from plasma-induced oxygen vacancies (Vo•) and Sr adsorption, which strengthen the metal-support interaction between Ir NPs and TiO2. Vo• and Sr promote charge transfer from TiO2 to Ir, optimizing the adsorption of OER/HER intermediates on oxygen- and hydroxyl-covered Ir NPs. Strong MSI also enhances structural stability against chemical corrosion. It is worth noting that the OER activity of the Ir@Sr-p-TiO2 NWs is evaluated by Sr adsorbed d-TiO2(004) with presence of an Ir cluster of 12 atoms (Ir@Sr-d-TiO2(004), showed in Figure 9a. Figure 9b shows the OER processes on Ir@Sr-d-TiO2(004) including four steps. Surprisingly, the catalyst requires overpotentials of only 250 and 32 mV to drive 10 mA cm−2 for OER and HER, respectively, surpassing commercial IrO2 and Pt/C. This design underscores the importance of defect engineering and MSI in bifunctional catalysis.
Metal–organic frameworks offer tunable platforms for integrating single atoms with defect-rich supports, enabling covalently electronic metal–support interactions (EMSI) that accelerate interfacial charge transfer. Ren et al. engineered Ru single atoms on defect-rich UiO-66 (Zr) via photochemical strategies. Covalent-bond-mediated EMSI between Ru SAs and UiO-66 accelerates interfacial charge transfer, enhancing photocatalytic activity [35]. The introduction of defects increased the NH3 yield from 4.57 to 16.28 μmol g−1 h−1, while Ru loading further boosted it to 53.28 μmol g−1 h−1. DFT studies reveal that EMSI-induced charge redistribution directs Ru d-orbital electrons into N2 π*-antibonding orbitals, activating the N≡N bond. This work advances MOF-based photocatalysts and deepens mechanistic understanding of N2 activation.
The nature of the EMSI profoundly influences the electronic configuration of single-atom catalysts (SACs), particularly through orbital-level metal-support interactions that dictate adsorption behavior and reaction pathways. Zhang et al. investigated Cu single-atom catalysts on metal oxides (Al2O3, CeO2, TiO2) to correlate EMSI with CO2RR performance [36]. EMSI-induced charge transfer modulates Cu’s electron density, yielding distinct highest occupied orbitals: Al2O3-CuSAC: 3dyz facilitates CO adsorption via 3dyz-π* backdonation, weakening C-O bonds and lowering C–C coupling barriers for multi-carbon products. TiO2-CuSAC: 3dz2 hybridizes with H2O σ/σ* orbitals, over-activating H2O dissociation to favor HER over CH4. CeO2-CuSAC: 3dx2−y2 aids CO2 activation, while localized Cu states suppress C–C coupling. Moderate H2O activation enables *CO hydrogenation with controlled HER. This study provides a comprehensive orbital-level framework for tailoring SAC-support combinations to achieve product-selective CO2RR.
Singhvi et al. designed Ni-Cu-Zn/CeO2 catalysts with strong SMSI between trimetallic nanoparticles and defective CeO2 [37]. XANES and XAS analyses confirm the electron redistribution from Ni 4s/4p to CeO2 which forms SMSI, while Cu and Zn develop partial positive charges via Ce interactions. Cu-Ni alloying boosts CO selectivity and stability, while Zn doping enhances CO adsorption and tunes Ni’s electron density. The catalyst achieves 49,279 mmol CO g−1 h−1 at 650 °C with 99% selectivity and 100 h stability, highlighting the synergy of electronic modulation and defect dynamics in CO2 utilization.

2.4. Modulation of Spin States and Magnetic Coupling

2.4.1. Impact of Spin-State Transitions on Catalytic Reactions

Spin state modulation enables the regulation of electronic structures, modifies the hybridization of electron orbitals, and elevates the d-band center. By altering the spin configuration of transition metal centers, we can reorganize orbital occupancy and facilitate the adsorption and activation of intermediates, thereby enhancing catalytic efficiency.
In regard to this strategy, Wang et al. introduced multiple defects to induce coordination unsaturation at Fe sites in NiFe-layered double hydroxide, increasing d-orbital splitting energy and stabilizing low-spin Fe3+ (LS Fe3+, t2g5eg0, Figure 10) [38]. XANES analysis revealed a disordered NiFe-LS structure due to high defect density, featuring electron transitions from 1s to unoccupied 3d orbitals. Spatially resolved XAS demonstrated spin-dependent electron reorganization between t2g and eg orbitals, altering hybridization between metal 3d and O 2p states. Compared to high-spin Fe3+ (HS, t2g3eg2), the LS configuration weakened O* intermediate adsorption at Fe3+ sites and established an optimal adsorption hierarchy (OH* > O* > OOH*), shifting the rate-determining step from O* → OOH* to more energetically favorable OH* → O*. The catalyst achieved an ultralow overpotential of 244 mV at 500 mA cm−2 (110 mV lower than HS NiFe-LDH), outperforming most reported NiFe-based systems. This work establishes a mechanistic link between spin-state transitions and adsorption energetics, positioning spin modulation as a key axis in rational catalyst design.
The co-engineering of lattice-site defects and oxygen vacancies offers a synergistic pathway to modulate spin states and electronic configurations of catalysts. Pei et al. co-incorporated La-site defects and oxygen vacancies into MOF-derived La0.8FeO3−δ, activating peroxymonosulfate (PMS) for levofloxacin (LVX) degradation under visible light [39]. Dual defects elongated Fe–O bonds, triggering a spin-state transition from low-spin to medium-spin Fe, which upshifted the d-band center and reduced the Fe(II)-PMS* activation barrier. This promoted Fe(IV)=O formation, enhancing PMS adsorption and LVX degradation. The catalyst exhibited exceptional stability under high salinity and natural organic matter, which photoelectrons promote the reduction of Fe(III) to Fe(II), thereby achieving the complete cycle of Fe(II)-Fe(IV)-Fe (II). This study exemplifies how defect-driven spin-state modulation can strategically enhance Fenton-like processes, extending its applicability in advanced oxidation technologies.
Huang et al. demonstrated chemically controlled spin-state transitions in a {Fe-Pt} Hofmann clathrate [40]. High-spin (HS) Fe(II) only facilitated photocatalytic ORR, requiring sacrificial agents. Low-spin (LS) Fe(II) enabled simultaneous ORR and water oxidation (WOR), achieving a H2O2 production rate of 66,000 μM g−1 h−1 under visible light. In situ characterization and DFT revealed LS-induced charge transfer between Fe(II) and Pt(IV), lowering energy barriers for ORR/WOR. Beyond water splitting, the system was further applied in upcycling sodium alkenesulfonate to high-value bromopropane, showcasing spin-state gating of cascade reactions. This work pioneers the use of spin-crossover modulation for programmable photocatalysis, showcasing how electronic spin modulation can couple reaction selectivity with catalytic functionality.

2.4.2. Defect-Engineered Magnetic Synergy: Tailoring Cooperative Spin Interactions via Structural Imperfections

Defect engineering induces interatomic ferromagnetic coupling, which imparts a spin-polarized magnetic response to the material, thereby enhancing its catalytic performance under external magnetic fields. This spin polarization facilitates spin-selective charge transfer and intermediate adsorption, particularly beneficial for spin-sensitive catalytic processes such as oxygen evolution or peroxide formation. Sun et al. developed a platform for synthesizing single-atom spin catalysts with broadly tunable magnetic substituents (M1) in MoS2 matrices [41]. Among all M1/MoS2 systems, Ni1/MoS2 adopts a distorted tetrahedral configuration, enabling ferromagnetic coupling between Ni1 sites and proximal S atoms. This induces global room-temperature ferromagnetism, which facilitates spin-selective charge transfer to generate triplet O2 during the OER. In water-splitting cells, Ni1/MoS2 exhibits exceptional magneto-catalytic performance, delivering a current density ≈10-fold higher (1.5–1.8 V) under a 502 mT field compared to zero-field conditions, outperforming commercial IrO2. The enhancement arises from field-induced spin alignment and optimized spin density distribution at S active sites, driven by field-modulated Ni–S orbital hybridization. This tailors adsorption/desorption energetics for radical intermediates (*OH/*O), lowering the O2 formation barrier by 0.072 eV in ferromagnetic configurations versus antiferromagnetic counterparts. Ni1/MoS2 also demonstrates superior stability in pure and seawater electrolytes, highlighting its potential for practical brine electrolysis. These findings establish ferromagnetic single-atom catalysts as platforms for spin-polarized electrocatalysis.
Engineering cation vacancies offer a powerful strategy for modulating spin states and surface electronic structures, with profound implications for magnetic ordering and catalytic activity. Wang et al. synthesized β-Ni(OH)2 nanosheets with tunable Ni2+ vacancy concentrations via hydrothermal growth followed by alkaline etching. The introduction of Ni2+ vacancies trigger a spin-state transition, inducing orbital charge redistribution and asymmetric surface charge density [42]. This vacancy-driven modulation converts Ni(OH)2 from antiferromagnetic to ferromagnetic, while activating electronic states near the Fermi level to enhance intrinsic OER activity. Under magnetic fields, vacancy-rich Ni(OH)2 exhibits pronounced spin-electron exchange with oxygen intermediates, reducing the OER overpotential by 20 mV at 20 mA cm−2. In contrast, pristine β-Ni(OH)2 shows negligible magnetic response. The synergy between defect engineering and spin manipulation offers a paradigm for designing magneto-responsive catalysts, with implications for spin-level mechanistic insights in electrocatalysis.

2.5. Dynamic Defect and Interface Engineering

2.5.1. Defect Reconfiguration Under Reaction Conditions

The generation of dynamic defect sites via electrochemical or allied reactions enables precise modulation of a catalyst’s electronic structure under working conditions, thereby enhancing its intrinsic activity and operational durability. Ren et al. synthesized a self-supported coral-like Ce2W2O9 array on carbon cloth via a facile ion-exchange method [12], followed by in situ electrochemical surface reconstruction to induce W vacancies, which dynamically modifies the local bonding environment of Ce atoms. Experimental and theoretical analyses reveal that optimized Ce sites lower the energy barrier of the potential-determining step (*OOH formation), accelerating O2 generation and enhancing intrinsic activity. In Ce2W2O9, Ce injects electrons into W sites via Ce–O–W chains. W vacancy formation during reconstruction activates adjacent Ce centers, tuning the Ce–O bond strength and optimizing oxygen intermediate adsorption energetics. The reconstructed R- Ce2W2O9 electrode achieves exceptional stability and durability, delivering 351 mV overpotential at 100 mA cm−2 and maintaining performance for >1000 h at 20 mA cm−2. This work provides mechanistic insights into rare-earth oxide electrocatalysts for industrial OER applications.
Controlling interfacial atomic mobility represents a promising strategy for enhancing catalyst stability and activity, particularly in alkaline HER systems. Li et al. engineered a nitrogen-doped carbon shell-encapsulated NiCoS heterostructure (CN@NiCoS) featuring dynamically regulated sulfur migration [43]. Sulfur vacancies at the Ni3S2–Co9S8 interface served as migration channels, and the migrating S atoms are captured via C–S bonds in the CN shell, inhibiting sulfide dissolution. The synergistic effect of sulfur-doped CN and S-vacancy pairs induces an upshift in the d-band center near the Fermi level, thereby enhancing HER kinetics and durability. The catalyst exhibits low overpotentials of 4.6 mV (freshwater) and 8 mV (seawater) at 10 mA cm−2, alongside 1000 h stability. This strategy guides interfacial atomic mobility design for robust alkaline HER.
Wang et al. anchored RuO2 nanoparticles on LiCoO2 nanosheets (RuO2/LiCoO2) to construct a catalyst capable of dynamic self-optimizing during OER [44]. The electron-donating LiCoO2 induces charge transfer from Co to Ru, resulting in stabilized Ru valence. Operando Li+ diffusion and extraction trigger interfacial restructuring, moderately weakening Ru–O covalency to suppress lattice oxygen participation while balancing activity and stability. This restructuring optimizes Ru sites for *OOH intermediate adsorption, thereby reducing the rate-determining step barrier. The catalyst achieves 10 mA cm−2 at 150 ± 2 mV overpotential with >2300 h stability in 0.5 M H2SO4 and operates continuously for 2000 h in proton exchange membrane electrolyzers at 1 A cm−2. This work resolves the stability-activity trade-off in acidic OER catalysts.

2.5.2. Heterojunction Interface Engineering

The construction of heterojunction interfaces induces an interfacial electric field, which significantly enhances charge transfer and redistribution at the interface. This built-in electric field facilitates the directional migration of photogenerated carriers and promotes the adsorption of reaction intermediates, thereby optimizing catalytic efficiency.
Engineering interfacial electric fields within heterostructures have proven highly effective for regulating electronic structure and accelerating multistep reaction kinetics. As an important heterogeneous component which can be applied to different properties [45,46], WO3 can be combined with other materials to construct the interfacial electric field and thereby regulate the electronic state. Wang et al. constructed a robust interfacial electric field in WS2–WO3 heterostructures, significantly accelerating the electrochemical nitrogen reduction reaction kinetics [11]. Schematic diagram of WS2–WO3 synthesis and reaction on the surface is shown in Figure 11. The abundant heterogeneous interfaces in WS2–WO3 possess strong interfacial electric fields that upshift the d-band center of W atoms, enhancing the adsorption of intermediate species (*–NH2 and *–NH). This interface enables lower energy barriers for *–NH2 and *–NH adsorption compared to pristine components, thereby shifting the rate-determining step from *NNH formation to *N generation while reducing overall surface energy barriers. The optimized WS2–WO3 heterostructure demonstrates superior catalytic performance with an NH3 production rate of 62.38 μg h−1 mg cat−1 and a Faradaic efficiency of 24.24%. This work elucidates how interfacial electric fields in WS2–WO3 regulate d-band centers and intermediate adsorption to enhance ENRR, providing critical insights for rational catalyst design.
Defect-engineered heterojunctions incorporating oxygen vacancies offer a powerful strategy to modulate interfacial charge distribution and promote lattice oxygen-mediated mechanisms in OER catalysis. Huang et al. developed a defect-engineered Fe2O3@CeO2-OV heterojunction with abundant oxygen vacancies supported on nickel foam for efficient OER [47]. Oxygen vacancies predominantly localized on CeO2 nanocluster surfaces intensify Fe2O3-CeO2 interactions, inducing strong CeO2 → Fe2O3 charge transfer and interfacial electron density redistribution. These OVs promote robust cluster-support interactions, downshifting the Ce-5d band center relative to Fe-3d orbitals, thereby reducing free energy barriers for *(OL-O) formation and O2 desorption steps in the lattice oxygen-mediated mechanism. The Fe2O3@CeO2-OV electrode achieves ultralow overpotentials of 172 mV and 317 mV at 10 mA cm−2 and 1000 mA cm−2, respectively, with exceptional stability.
Coupling plasmonic materials with strain- and defect-engineered semiconductors provides a promising route for enhancing light harvesting and catalytic activation in solar-driven nitrogen fixation. Li et al. designed a CeO2-AD/Au heterostructure by anchoring strain-engineered CeO2 nanosheets with Vo defects onto Au hollow nano-mushrooms (HNMs) for N2 photofixation [48]. Plasmonic Au HNMs broaden light absorption to the NIR-II region while generating hot electrons/holes. Atomically sharp CeO2/Au interfaces facilitate rapid hot electron transfer, enhancing utilization efficiency. Concurrently, interface compressive strain induces abundant strain-OV defects that promote N2 adsorption and in situ activation. This synergistic structural/defect engineering enables solar-driven nitrogen fixation efficiency comparable to natural photosynthesis, highlighting its potential for artificial photosynthetic systems.
Integrating oxygen vacancies into heterojunctions offers a dual advantage: modulating interfacial electronic configurations and enhancing carrier transport dynamics. Zu et al. synthesized WO3-OV/In2S3 S-scheme heterojunctions through hydrogen-thermal treatment, where OVs alter interfacial electronic configurations [49]. Sulfur-anchored OVs increase electron density near W 5d band, forming intimate interfacial contact that enhances charge transfer. Femtosecond transient absorption spectroscopy (fs-TAS) reveals OV-induced electron trapping states enable additional charge transport pathways, improving charge separation efficiency by 34.5-fold and 2.9-fold compared to WO3-OV and In2S3 alone, respectively. The optimized heterojunction outperforms non-defective WO3/In2S3, demonstrating OVs’ dual role in carrier dynamics modulation and thermodynamic parameter optimization for photocatalysis.
S-scheme heterojunctions offer a strategic architecture for achieving both directional charge separation and selective retention of high-energy carriers, crucial for efficient photocatalysis. Deng et al. fabricated In2O3/Nb2O5 S-scheme heterojunctions, achieving sub-10 ps ultrafast interfacial electron transfer [50]. Fs-TAS confirms selective charge recombination between weak electrons (In2O3 CB) and holes (Nb2O5 VB), while preserving strong electrons (Nb2O5 CB) and holes (In2O3 VB) for photoreactions. CO2-TPD and DFT calculations verify enhanced CO2 chemisorption/activation at heterointerfaces. The optimized nanofibers achieve a CO production rate of 0.21 mmol g−1 h−1 without sacrificial agents, showcasing fs-TAS’s utility in probing S-scheme charge dynamics. This work establishes a framework for designing advanced heterostructures with dual-enhanced charge separation and reactant activation capabilities.

3. Multiscale Characterization of Defect–Electronic Structure–Performance Correlations

3.1. Defects Characterization

3.1.1. Electron Paramagnetic Resonance (EPR)

Electron paramagnetic resonance is a widely used technique for characterizing material defects. Its working principle relies on the Zeeman splitting of unpaired electron spin states under an external magnetic field, where specific microwave frequencies induce spin-state transitions. The resulting absorption signals provide critical insights into electronic spin configurations. EPR is particularly effective in probing transient radical intermediates in chemical reactions, determining oxidation states and coordination geometries of transition metal ions, and investigating semiconductor defects (e.g., dopant sites) and catalytic active centers. Through its ability to resolve spin-sensitive interactions at the atomic scale, EPR has become an indispensable tool for elucidating structure–activity relationships in catalytic and electronic materials.
In the study by Wu et al., defect-rich iron-nickel oxyhydroxide (d-(Fe, Ni)OOH) was synthesized for efficient OER [51]. The introduction of engineered oxygen vacancies not only enhanced active site exposure and maximized the catalyst’s effective surface area but also modulated the local coordination environments and chemisorption properties of Fe/Ni sites. EPR confirmed the oxygen vacancies, with d-(Fe, Ni)OOH exhibiting a stronger magnetic signal at g ≈ 2.004 compared to pristine (Fe, Ni)OOH, indicating a higher unpaired electron density induced by oxygen defects (Figure 12).
In MOFs-related materials, Ding et al. developed defective MOFs (denoted as NiFc*′xFc1–x) featuring unsaturated metal sites [52]. These defects facilitated in situ transformation into metal hydroxides during OER, yielding materials with abundant oxygen vacancies that promoted the adsorption of oxygen intermediates and significantly boosted OER activity. EPR analysis revealed an intensified and broadened characteristic signal at g = 2.13 for NiFc′Fc, attributed to increased unpaired electrons from Fc incorporation.
In recent studies about defects in transition metal oxides, Bi et al. demonstrated the synthesis of Ni-doped TiO2 containing triply associated O-Ti-O vacancies (VOVTiVO) [53]. Hydroxyl groups adsorbed on unsaturated cationic sites critically promoted VOVTiVO formation by facilitating H2O dissociation and OH* deprotonation, enabling defect regeneration. EPR spectra showed distinct signals at g = 2.002 (unpaired electrons at oxygen vacancies). Ni-TiO2-N2 exhibited the strongest signal, confirming abundant oxygen vacancies. Strikingly, Ni-TiO2-Air displayed an additional signal at g = 1.99, attributed to Ti vacancies, and g = 1.97 (Ti3+ centers in TiO2), highlighting annealing atmosphere dependent vacancy configurations.

3.1.2. Positron Annihilation Lifetime Spectroscopy (PALS)

Positron Annihilation Lifetime Spectroscopy is a widely used technique for investigating material defects. When high-energy positrons (β+) are implanted into a material, they annihilate with electrons, emitting gamma rays. By measuring annihilation lifetimes, the size and concentration of defects such as vacancies and voids can be deduced. PALS enables quantitative evaluation of microscopic defects in metals, ceramics, polymers, and other materials. In addition, PALS is also employed to track crystallization processes in amorphous materials, segmental motions in polymer chains, pore size distribution/connectivity in nanoporous materials (e.g., MOFs, zeolites), and radiation-induced defects in nuclear or semiconductor systems.
In-depth characterization of vacancy configurations in doped metal oxides can be achieved through a combination of PALS and DFT simulations. In the study by Bi et al., triple O-Ti-O vacancy associates (VO-VTi-VO) in Ni-doped TiO2 nanoparticles were characterized using PALS combined with DFT calculations [53]. All samples exhibited three distinct lifetime components (τ1, τ2, and τ3) in their PALS profiles: τ1 (~250 ps): Related to positron annihilation in defect-free regions or small defects (e.g., single vacancies or vacancy clusters). τ2 (350–550 ps): Corresponded to large vacancy clusters in crystalline and grain boundary regions. τ3 (5–15 ns): Attributed to positronium annihilation at large voids or interfaces. DFT calculations revealed that the primary defect in Ni-TiO2-air (τ1 = 232.3 ± 4.1 ps) originated from O-Ti-O trimeric vacancy associates (VO-VTi-VO), while Ni-TiO2 (197.8 ± 2.2 ps) and Ni-TiO2-N2 (197.7 ± 4.5 ps) contained isolated single vacancies (VO, VTi) or O-O vacancy clusters (VO-VO/VO-VO-VO), with predicted lifetimes of 189.8–199.0 ps. This combined experimental–computational framework enables atomic-level identification of vacancy types and configurations, offering a robust strategy for elucidating defect structures in functional metal oxides.
Recent studies about multicomponent nanozyme systems also showed the effectiveness of PALS. Zhang et al. developed a symbiotic nanozyme system by anchoring bacteroid-like CeOx nanoclusters onto Mn3O4 nano-supports (CeOx/Mn3O4) [54]. A “material exchange” process between Ce and Mn atoms optimized the ratios of Ce3+/Ce4+ and Mn3+/Mn2+ while enhancing vacancy concentrations via interfacial defect engineering. The resulting CeOx/Mn3O4 exhibited robust catalase-like (CAT-like) and superoxide dismutase-like (SOD-like) activities. PALS analysis gradually increased τ1 values (163.9 ± 7.3 ps for CeO2, 167.1 ± 1.3 ps for Mn3O4, and 204.3 ± 2.6 ps for CeOx/Mn3O4), indicating a rise in both the number and size of small vacancy-type defects within the annealed symbiotic system (Figure 13). The τ1 ≈ 163 ps signal in CeO2 was attributed to VO’’ and VO’’VO’’ vacancy configurations.

3.2. Electronic Structure Characterization

3.2.1. X-Ray Photoelectron Spectroscopy (XPS) and Ultraviolet Photoelectron Spectroscopy (UPS)

X-ray Photoelectron Spectroscopy and Ultraviolet Photoelectron Spectroscopy are among the most widely used techniques for analyzing the electronic structures of nanomaterials. XPS, a surface-sensitive analytical method based on the photoelectric effect, provides critical insights into elemental composition, chemical states, and electronic configurations at material surfaces. It enables the direct identification of surface elements, quantitative concentration analysis, differentiation of oxidation states, and even functional group characterization. In contrast, UPS employs ultraviolet photons to excite valence electrons, elucidating valence band energy-level distributions. This technique is particularly suited for studying organic compounds, semiconductors, and adsorbed species, as it measures work functions, determines band structures (e.g., direct/indirect bandgaps in semiconductors/insulators), and probes molecular orbitals such as the highest occupied molecular orbital (HOMO) and electron delocalization. Together, XPS and UPS offer complementary insights into surface chemistry and electronic structure, making them indispensable tools in the characterization of nanoscale materials and interfaces.
For intermetallic alloy catalysts, XPS successfully determines the surface characteristics including engineered porosity and defect architectures Wu et al. synthesized a highly porous intermetallic alloy (CoCuMoNi) with abundant defect sites [55]. The CoCuMoNi catalyst possesses the bifunctional water decomposition performance of both HER and OER, attributed to the ternary interaction of heterogeneous structure interfaces, abundant defects and internal electric fields. XPS analysis revealed the chemical states and electron transfer dynamics in the nanoporous CoCuMoNi electrode and control samples (Figure 14). Multiple oxidation states were identified: Ni0 and Ni2+ for Ni; Mo0, Mo4+, and Mo6+ for Mo; Cu0, Cu1+, and Cu2+ for Cu; Co0, Co2+, and Co3+ for Co. The electronegativity differences among the elements and the heterojunction structure induced significant interfacial charge redistribution. Covalent bonding between Mo-Ni, Cu-Ni, and Co-Ni facilitated electron transfer from less electronegative Mo to more electronegative Ni, followed by further transfer to Cu and Co. Conversely, Co-Mo and Co-Cu bonding enabled electron transfer from Mo to Co and subsequently to Cu. Overall, Ni and Mo acted as electron donors, while Cu and Co served as acceptors, forming a robust, highly active catalytic architecture.
In terms of intra-band defect levels, Xu et al. developed a WO3−x/In2S3 S-scheme heterojunction with tunable intra-band defect levels for enhanced CO2 photoreduction [13]. The charge transfer pathway was investigated via radical trapping experiments and in situ irradiated XPS (ISI-XPS). Compared to pristine WO3−x and In2S3, the composite showed significantly intensified DMPO-•OH and DMPO-•O2 signals, confirming efficient separation and preservation of high-energy electrons (In2S3 conduction band, CB) and holes (WO3−x valence band, VB). Under illumination, W 4f and O 1s binding energies (BEs) in WO3−x/In2S3 shifted positively, while In 3d and S 2p BEs exhibited negative shifts, directly evidencing electron transfer from WO3−x to In2S3 and validating the S-scheme mechanism. Notably, WO3−x/In2S3-550 displayed the most pronounced BE shifts, indicating that optimized intra-band defect levels enhanced charge carrier transfer efficiency.
Li et al. engineered a TiO2 heterostructure, c-TiO2@a-TiO2−x(OH)y, featuring surface-frustrated Lewis pairs of HO-Ti-[O]-Ti embedded in an amorphous shell surrounding a crystalline core [56]. These SFLPs can heterolytically dissociate H2, forming protonated hydroxyls and hydrides, which achieved CO2 reduction. UPS analysis mapped the band-edge positions: the valence band edges of c-TiO2 and a-TiO2−x(OH)y (0.0015 < x < 0.0031) were −7.7 eV and −7.59 eV (vs. vacuum), respectively, with corresponding conduction band edges at −4.13 eV and −4.21 eV (Figure 15). The band structure information obtained from UPS provides a basis for analyzing the electron-hole transfer, separation and recombination mechanisms.

3.2.2. X-Ray Fine Structure Analysis (XAFS)

X-ray Absorption Fine Structure is an advanced characterization technique that analyzes subtle energy-dependent variations in X-ray absorption. It provides quantitative insights into the local atomic structure including coordination environment, bond lengths, and chemical states. It comprises two core components: X-ray Absorption Near-Edge Structure (XANES) and Extended X-ray Absorption Fine Structure (EXAFS). XANES focuses on the fine structure near the absorption edge, where edge shifts reflect oxidation states and pre-edge peak intensity/shape reveals orbital hybridization. EXAFS analyzes post-edge oscillations, using frequency to determine bond lengths, amplitude to quantify coordination numbers, and phase shifts to identify neighboring atoms.
Yang et al. engineered Fe–N3 sites with varying Fe–N bond lengths via carbon defect introduction, leveraging XAFS for structural differentiation [57]. Fe K-edge XANES spectra exhibited similar pre-edge features and line shapes due to shared “T” geometry across Fe–N3 configurations, yet the main peak shifted systematically (Figure 16). Compared to the reference (M0), M1 and M4 showed lower main peak energies, while M3, M5, and M6 displayed higher energies. According to Natoli’s rule, these shifts inversely correlate with the average Fe–N bond length squared. By modifying bonding distances via defects, the local coordination structure of Fe–N3 sites was tailored, directly impacting catalytic thermochemical activity. Critically, this structural tuning was quantitatively linked to XANES features, establishing an experimentally accessible structure–activity descriptor.
X-ray analyses are also capable of determining mixed metal valence state, which is crucial for transition metals. Liu et al. designed a heteronuclear FeMn dual-atom catalyst anchored on nitrogen-doped carbon (FeMnDSA/dNC) using a defect-trapping strategy [58]. This defect engineering optimized electronic structures toward the Sabatier optimum, significantly enhancing ORR performance. XAFS elucidated the electronic/coordination environments: Fe K-edge XANES positioned absorption edges between FeO and Fe2O3, indicating mixed Fe2+/Fe3+ states. The k3-weighted Fourier-transform EXAFS revealed a primary Fe–N shell at ~1.50 Å and a heterometallic Fe–Mn scattering path at ~2.45 Å. Mn K-edge XANES matched Mn phthalocyanine (Mn2+), with EXAFS confirming Mn–N bonds (~1.38 Å) and Fe–Mn distances (~2.45 Å). EXAFS fitting and wavelet transform analysis validated the heteronuclear Fe–Mn pairs: Fe coordinates with four N atoms and one Mn atom, while Mn adopts a similar N4 configuration with Fe coupling. Distinct WT intensity maxima (Fe: 4.8 Å−1; Mn: 5.1 Å−1) diverged from metallic Fe/Mn standards (8.2 and 7.7 Å−1), confirming atomic dispersion.
In addition, Zhou et al. developed a defect-driven activation strategy for Co-based MOFs (D/CoFc-MOF), tuning adsorption behavior via local coordination and electronic reconfiguration [59]. XAS analysis revealed that Co K-edge XANES showed edge shifts toward higher energy in D/CoFc-MOF vs. CoFc-MOF, suggesting increased Co oxidation states (+2.1 to +2.2). EXAFS identified the first Co–O coordination shell (~1.5 Å), with shortened bond lengths in D/CoFc-MOF indicating enhanced Co–O covalency (Figure 17). WT-EXAFS at ~3.8 Å−1 confirmed Co–O paths without Co–Co contributions, proving atomic dispersion. Electronic structure analysis via orbital-projected partial density of states (PDOS) revealed t2g (dxy, dyz, dxz) and eg (dx2, dz2) orbital occupations, with defect engineering reducing d-electron counts (7.15 → 7.13 e) at Co orbitals and increasing the oxidation state of Co.

3.2.3. Electron Energy Loss Spectroscopy (EELS)

Electron Energy Loss Spectroscopy is a characterization technique that analyzes energy losses of high-energy electrons interacting with materials to determine the composition, electronic structure, and chemical environments, often coupled with transmission electron microscopy (TEM). EELS enables elemental identification via ionization edges, resolves unoccupied electronic states through near-edge fine structures, distinguishes chemical states, and extracts coordination information via extended energy loss fine structure (EXELFS) oscillations. Additionally, EELS probes valence/conduction band structures and interfacial charge transfer dynamics.
Li et al. anchored strain-engineered CeO2 nanosheets with abundant strain-vo defects onto Au hollow nano-mushrooms, constructing a semiconductor/plasmonic metal heterostructure [48]. Interfacial compressive strain induced strain-vo defects that enhanced nitrogen adsorption and in situ activation. Aberration-corrected HAADF-STEM and EELS mapped Ce3+ and Ce4+ distributions in CeO2-AD/Au (Figure 18). EELS spectra of Ce M-edges revealed reduced peak positions in regions near Au HNMs (Region 1: M5-edge by 1.8 eV, M4-edge by 2.0 eV) compared to bulk CeO2 (Region 2). The M5/M4 intensity ratio in Region 1 matched Ce3+, while Region 2 aligned with Ce4+. Oxygen K-edge EELS highlighted weakened peak a intensity (O–Ce4+ bonds) in Region 1, corroborating Ce3+ predominance at the CeO2/Au interface. This Ce3+ enrichment, stabilized by strain-induced vo defects, enhanced catalytic stability. In contrast, CeO2-LD/Au exhibited significantly reduced interfacial Ce3+ content, as quantified by EELS valence mapping.
In another study, Li et al. synthesized a TiO2 heterostructure, c-TiO2@a-TiO2−x(OH)y, featuring surface-frustrated Lewis pairs of HO-Ti-[O]-Ti in an amorphous shell encapsulating a crystalline core [56]. Ti L-edge EELS at interfacial regions revealed structural and electronic disparities: Core regions showed characteristic Ti(IV) L2,3-edge splitting, consistent with crystalline TiO2 symmetry. Surface regions exhibited suppressed splitting and a −2.2 eV energy shift, indicative of Ti3+-rich amorphous a-TiO2−x(OH)y with distorted Ti–O6 octahedra and O vacancies. The reduced t2g-e2g splitting and symmetry breaking at the interface were attributed to charge redistribution and stabilization of low-valent Ti in oxygen-deficient amorphous domains.

3.3. Theoretical Calculations

3.3.1. Density Functional Theory (DFT) Calculations

Density Functional Theory is a quantum mechanics-based computational method that investigates the ground-state properties of many-body systems by describing electron density rather than wavefunctions. As one of the most widely used theoretical tools in computational chemistry and materials science, DFT enables predictions of energy, structure, electronic properties, and reaction mechanisms for molecular, crystalline, and surface systems. Its applications span from evaluating molecular stability assessments, calculating band structure and density of states, mapping charge distribution, and analyzing transition-state/reactivity.
In the study by Zhou et al., DFT calculation was applied to analyze the electron properties of MOFs. They developed a defect-driven activation strategy to control adsorption behavior in Co-based MOFs (denoted as D/CoFc-MOF) by modulating local coordination geometry and electronic configurations [59]. DFT calculations were performed to analyze DOS, crystal orbital Hamilton population (COHP) of Co–O bonds, d-band center shifts (Ed), and energy-level alignments. Structural models included a pristine CoFc-MOF and its defect-engineered variant (D/CoFc-MOF) with non-bridging Fc’ ligands, where Co sites were preferentially exposed as active centers for oxygen evolution reaction (OER). The defect-induced band engineering markedly changed the Ed level from −1.60 eV (CoFc-MOF) to −1.83 eV (D/CoFc-MOF), optimizing electron transfer efficiency and reshaping electron density distributions. This enabled balanced adsorption/desorption of oxygen intermediates (OH*, O*, OOH*) through enhanced Co–O covalency, as evidenced by: Stronger orbital overlap between Co-3d and O-2p states in D/CoFc-MOF. Integrated COHP (ICOHP) values confirm tighter Co–O bonding compared to CoFc-MOF.
In another study, Chen et al. utilized ultrasonic cavitation to introduce controlled defects into FeCoNi/FeAl2O4 hybrid coatings, optimizing OER activity [60]. DFT studies on the heterojunction’s electronic structure employed: Special quasi-random structure (SQS) models for FeCoNi (metallic phase) and FeAl2O4 (spinel phase), validated against HAADF-STEM data. Charge density difference maps showing electron accumulation at the heterointerface, with Bader charge analysis confirming electron depletion in FeCoNi and accumulation in FeAl2O4. Projected DOS (PDOS) revealing d-d orbital coupling between Co/Ni 3d states near EF, critical for intermediate stabilization. Broad eg-t2g splitting in Fe 3d orbitals, indicative of rapid electron transfer, and Al 2p/O 2p hybridization enhancing metal–oxygen covalency. These computational insights rationalized the superior OER performance, with Ed shifts and interfacial charge redistribution serving as key descriptors for defect-mediated catalyst optimization.

3.3.2. Machine Learning(ML)-Driven Prediction

Machine Learning leverages data-driven pattern recognition to predict chemical compositions, properties, reaction pathways, and material performance, significantly accelerating chemical research and materials discovery. Lin et al. demonstrated a closed-loop workflow integrating ML-assisted prediction, multi-objective optimization, and experimental synthesis to coordinately optimize defect content and thermal stability in UiO-66(Ce), achieving enhanced catalytic hydrogenation of dicyclopentadiene [61]. The ML workflow comprised three key steps:
  • Automated Data Extraction: A line-recognition algorithm and thermogravimetric analysis curve evaluation method were developed to automate data extraction and batch calculations, minimizing manual errors and improving efficiency.
  • Pure-Phase Subspace Delineation: Given the sensitivity of UiO-66(Ce) synthesis to parameters, ML screened >200,000 experimental protocols to define a tailored synthesis subspace, preventing impurity phases like CSUST-1 or Ce(HCO2)3.
  • Closed-Loop Optimization: A hybrid utility function combining probability of achievement and weighted-sum techniques balanced defect density and thermal stability.
This enabled Pareto-frontier optimization, achieving target MOFs (Td > 300 °C, defect content > 40%) within two experimental iterations. The optimized UiO-66(Ce) exhibited approximately 10-fold higher activity in DCPD hydrogenation compared to conventional approaches. This work establishes a universal data-driven paradigm for intelligent materials synthesis, advancing beyond traditional trial-and-error methods reliant on prior knowledge.

4. Defect Engineering-Driven High-Efficiency Catalytic Applications

4.1. Hydrogen Evolution Reaction (HER)

Hydrogen Evolution Reaction refers to the electrochemical process of generating hydrogen gas at the electrode surface via reduction, a pivotal reaction in water-splitting technologies for hydrogen production. As a key half-reaction in water-splitting and electrolysis systems, HER plays a central role in sustainable hydrogen production and renewable energy conversion technologies. Defect engineering has emerged as a powerful strategy to enhance HER activity by increasing active sites, modulating electronic structures, promoting charge transfer, and enabling surface reconstruction and synergistic effects.
Heteroatom doping is an effective strategy to manipulate the electronic structure and defect states in photocatalysts for enhanced hydrogen evolution performance. Pei et al. synthesized a boron-doped CRP (B-CRP) photocatalyst via a mild boric acid-assisted hydrothermal strategy. Boron incorporation filled vanadium vacancies, suppressed deep trap states, and introduced shallow trap states in CRP’s band structure [62]. These STS prolonged charge carrier lifetimes by inhibiting deep trapping and recombination, while boron dopants acted as active sites to boost HER. The optimized B-CRP achieved a photocatalytic HER rate of 1392 μmol g−1 h−1, outperforming pristine CRP (4-fold) and ARP (10-fold). Stability tests confirmed structural integrity and sustained activity over multiple cycles.
Ligand-assisted defect and phase engineering provides a versatile route for tuning interlayer structure, orbital interactions, and enhancing HER catalytic activities. Liu et al. developed a Co-doped 1T-MoS2 catalyst (Co-1T-MoS2-bpe) via ligand modulation using a CoMo-MOF precursor [63]. The multidentate 1,2-bis(4-pyridyl)ethane (bpe) ligand expanded interlayer spacing, exposed active sites, and accelerated water dissociation. DFT calculations revealed that π-conjugated ligand coordination weakened Mo 3d/S 2p orbital hybridization, enhancing S-2p/H-1s orbital overlap for optimized H adsorption. Co-1T-MoS2-bpe delivered an ultralow overpotential of 118 mV at 10 mA cm2 in alkaline HER (Tafel slope: 83 mV dec−1) via the Volmer-Heyrovsky pathway, and achieved overpotential at 200 mA cm−2 which is lower than Pt/C. Integrated into an anion-exchange membrane water electrolyzer, it maintained 100 mA cm−2 at 2.0 V for 10 h, surpassing Pt/C performance.
In addition to metal doping, nonmetallic heteroatom doping also showed the capability for improving HER activities. Zhang et al. engineered N,B,P/S-codoped carbon catalysts (BPN-C, BSN-C) from ZIF-derived N-doped carbon [64]. Nonmetal dopants induced defects, redistributed electron density, and increased active sites. BSN-C exhibited superior HER activity in alkaline media: η10 = 129.7 mV and a Tafel slope of 72.7 mV dec−1, outperforming BPN-C, BN-C, and N-C. After 1000 cycles, BSN-C retained 90.68% of the onset overpotential, demonstrating exceptional stability.

4.2. Oxygen Evolution Reaction (OER)

The oxygen evolution reaction is an anodic process generating oxygen through oxidation, serving as a critical half-reaction in water splitting for oxygen production and a cornerstone of metal-air batteries and CO2 reduction technologies. Defect engineering—introducing vacancies, dopants, and grain boundaries—enhances OER activity by tailoring electronic structures and surface microenvironments to reduce reaction energy barriers.
Dynamic reconstruction of MOFs into active transition metal (oxy)hydroxides has emerged as a promising approach to OER activity through in situ defect and composition modulation. Zhang et al. incorporated monocarboxylic acid-functionalized linkers (ferrocene carboxylic acid, FcCA) into NiBDC-FcCA MOF, creating unsaturated sites that dynamically reconstruct into Fe-doped Ni(OH)2/NiOOH during OER [65]. Oxygen vacancies and Fe dopants persisted throughout reconstruction, optimizing electron transport and weakening oxygen intermediate adsorption on Ni sites to favor OOH* formation. NiBDC-FcCA exhibited a ninefold higher current density at 1.6 V vs. NiBDC. The optimized NiBDC-FcCA/Fe foam achieved an OER overpotential of 250 mV at 200 mA cm−2, with negligible activity loss after 1200 h at 1 A cm−2. When integrated into an anion-exchange membrane water electrolyzer, it delivered 1 A cm−2 at 1.85 V, outperforming commercial Pt/C||RuO2 (1.92 V).
Defect engineering in 2D MOFs offers a tunable platform for enhancing bifunctional electrocatalysis (OER/HER) by optimizing active site exposure and intermediate adsorption energetics. Zhao et al. employed targeted defect engineering to create high-density unsaturated Fe sites in 2D Fe-MOFs. These defects promoted H2O adsorption and *OH intermediate formation, modulating OER rate-determining steps [66]. Oxygen vacancies further optimized O and H adsorption, enabling dual OER/HER functionality. The optimized Fe-MOF achieved overpotentials of 259 mV (OER) and 36 mV (HER) at 10 mA cm−2 in 1.0 M KOH. Fe-MOF0.3 demonstrated exceptional stability in AEMWE under cyclic potentials (1.6–1.9 V), maintaining performance for 125 h.
Anion defect engineering offers a promising route to modulate lattice structure, charge distribution, and catalytic pathways for OER enhancement. Gu et al. introduced anionic iodine into Cu lattices, inducing charge redistribution and lattice distortion to regulate surface reconstruction during OER [67]. Iodine defects weakened Cu–S bonds, facilitating Cu2+ → Cu3+ oxidation and generating oxygen vacancies (VO). These VO sites provided spatial regions for OH adsorption and shifted the OER pathway to an oxygen vacancy oxidation mechanism (OVOM). The optimized VO-I0.033-CuOS achieved a 189 mV overpotential at 10 mA cm−2 and 1250 h stability. In an AEMWE electrolyzer, it delivered 1 A cm2 at 1.65 V with a hydrogen production cost of $1.70/kg, surpassing the U.S. DOE’s 2026 target.
Understanding the distinct roles of cationic and anionic vacancies is essential for rationally designing high-performance OER catalysts. Zhang et al. investigated the roles of O (VO) and Co (VCo) vacancies in defective Co3O4 in initial activity and reconstruction process [68]. Initial OER activity followed Co3O4-VCo > Co3O4-VO > pristine Co3O4, indicating the improvement from the introduction of Co or O vacancies. Specifically, VO enhanced *OH adsorption but slowed deprotonation, driving irreversible reconstruction into amorphous [Co(OH)6] intermediates, while Co vacancies accelerated deprotonation but reduced the reconstruction rate. Both O and Co vacancies yields reconstructed Co bridge sites with high OER activity, leading to the increase or decrease in the Co–Co bond length, respectively, which Co3O4-VCo has the best activity (η10 = 262 mV vs. 320/300 mV for pristine Co3O4/Co3O4-VO).

4.3. Carbon Dioxide Reduction(CO2RR)

The electrochemical and photochemical reduction of carbon dioxide offers a sustainable route to convert greenhouse gas into value-added fuels and chemicals. CO2 reduction is a catalytic process driven by external electric fields or light to convert CO2 into products including CO, CH4, C2H4, and HCOOH. Defect engineering—introducing vacancies, dopants, or grain boundaries—enhances CO2RR activity and selectivity by optimizing electronic structures, exposing active sites, promoting CO2 adsorption/activation, modulating intermediate adsorption energies, and improving charge transfer kinetics.
Engineering heterojunctions and anion vacancies provide a synergistic approach to enhance light absorption, charge separation, and intermediate stabilization in photocatalytic CO2 reduction. Lai et al. synthesized an In2S3/In2O3 spatial heterojunction (ISIO(VS)) with sulfur vacancies via polyvinylpyrrolidone treatment [69]. The heterojunction and S vacancies enhance visible-light utilization and charge separation. S vacancies lower the work function of In2S3, intensifying built-in electric fields to improve charge separation. DFT calculations reveal that S vacancies induce charge accumulation on adjacent In atoms, facilitating CO2 activation. In situ DRIFTS and ΔG analysis confirm that ISIO (VS) exhibits a lower energy barrier for *CHO formation than CO desorption, favoring CH4 selectivity (95.93%, 16.52 μmol g−1 h−1). S vacancies also strengthen In–C interactions in *CHO intermediates, lowering ΔG*CHO.
Gradient defect engineering enables spatial modulation of electronic structure and internal electric fields to enhance photocatalytic CO2 reduction. Wang et al. introduced a 3–4 nm gradient W vacancy layer on Bi2WO6 nanosheets, creating an inner-to-outer tandem homojunction with a graded Fermi level [70]. This directional internal electric field drives photoelectron migration from bulk to surface, providing a “highway” for charge separation. W vacancies alter the coordination of O/W atoms, shifting CO2 adsorption from weak/strong to moderate bonding, reducing the *COOH formation barrier. The defect-engineered Bi2WO6 achieves a CO production rate of 30.62 μmol g−1 h−1 with 99% selectivity, without requiring cocatalysts or sacrificial agents.
Defect engineering in layered double hydroxides (LDHs) offers an effective strategy to tune surface reactivity and adsorption properties for selective CO2 photoreduction. Wang et al. engineered oxygen vacancies (VO) on ZnAl LDH to regulate CO2 adsorption and surface reactivity [71]. VO alters the electronic structure of Zn/Al sites, enabling selective chemisorption of H2O and CO2. Under illumination, CO2 adsorbs on Al sites to form *COOH, which is subsequently protonated to *CO. Simultaneously, H2O dissociates at Zn sites to supply protons. Dynamic structural evolution (layer expansion and crystal contraction) enhances CO production, with Vo-ZnAl achieving 17.2 μmol g−1 h−1 significantly outperforming pristine ZnAl (6.3 μmol g−1 h−1).
Introducing linker vacancies in MOFs offers a strategic route to promote dynamic structural evolution during CO2RR catalysis. A et al. developed a Cu-MOF (UC-Cu-BTEC) with linker vacancies for CO2RR [72]. Vacancies induce distortion in Cu dimers within the framework, preserving structural integrity while creating undercoordinated Cu sites. During CO2RR, defective Cu dimers reconstruction generate abundant uncoordinated Cu species, enhancing *CO coverage and stabilizing C2 intermediates. This promotes C–C coupling, yielding a C2+ Faradaic efficiency (FEC2+) of 77.2% and sustaining FEC2+ > 70% across a 400 mV potential window. In flow cells, UC-Cu-BTEC achieves a C2+ partial current density of -153.5 mA cm−2 under neutral conditions.

4.4. Ammonia Synthesis Catalysis

Ammonia synthesis catalysis refers to the reduction of nitrogen sources (e.g., N2) into ammonia (NH3) via electrochemical or catalytic processes. This reaction is central to fertilizer production, energy storage, and emerging green hydrogen carriers. Defect engineering optimizes catalytic activity by modulating lattice defects (e.g., vacancies, doping, dislocations) to tailor electronic structures, enhance surface reactivity, suppress side reactions, and expose active sites for nitrogen activation.
Single-atom catalysts with engineered coordination environments offer a promising platform for high-efficiency nitrate reduction to ammonia. Ding et al. developed a defect-enriched coordination polymer (d-CoCP) with well-defined, unsaturated single-atom Co sites [73]. Methanol etching of a cobalt-based coordination polymer (CoCP) introduced nitrogen vacancies, creating localized undercoordinated Co sites. These vacancies reduced the energy barrier of the rate-limiting step in nitrate reduction to ammonia (NRA), enhancing NO3 adsorption and conversion. The optimized d-CoCP achieved an NH3 production rate of 0.78 mmol h−1 cm−2 at 173.3 mA cm−2 with 97% Faradaic efficiency (FE), surpassing pristine CoCP and other single-atom catalysts. Stability tests confirmed negligible activity loss over 100 h.
The coupling of intrinsic piezoelectric fields with defect engineering offers a unique strategy for enhancing nitrogen reduction reaction performance under piezo-photocatalytic conditions. Yuan et al. engineered BaTiO3 (BTO) with oxygen vacancies to synergize defects and piezoelectric fields for nitrogen reduction (NRR) [74]. Ti3+ sites near OVs possess unpaired d-orbital electrons, facilitating N2 activation. Piezoresponse force microscopy and Kelvin probe force microscopy revealed enhanced polarization and charge separation in Ti3+-OV-modified BTO. In situ characterization and DFT studies showed that piezoelectric fields shift the Ti3+ d-band center toward the Fermi level, lowering the energy barrier for the rate limiting step. The optimized BTO with moderate oxygen vacancies achieved an NH3 yield of 106.7 μmol g−1 h−1 under piezo-photocatalysis.
The size and spatial distribution of surface defects significantly influence nitrogen reduction performance by modulating adsorption dynamics and competing reaction pathways. Sun et al. synthesized VOx nanobelts with controlled defect sizes (0–1 nm, 1–2 nm, > 5 nm) [75]. Small defects limited spatial flexibility for N2 hydrogenation, while large defects favored competing HER. Medium-sized defects (1–2 nm) balanced electronic and geometric effects, promoting optimal N2 adsorption and activation. The intermediate defect size yielded an NH3 rate of 81.94 ± 1.45 μg h−1 mg−1 and FE of 31.97 ± 0.75% at −0.5 V vs. RHE, demonstrating the critical role of defect dimensions in eNRR efficiency.
The synergistic integration of surface-loaded noble metals and defect-rich transition metal frameworks provides a powerful strategy for enhancing nitrogen reduction electrocatalysis. Guo et al. fabricated multi-defect nanocrystalline Cu via thermal quenching, where stress-induced defects and nonequilibrium phase transitions enhanced intermediate adsorption and proton transport [76]. Surface-loaded Ir species provided active H*, synergizing with defect-rich Cu to modulate the d-band center and optimize intermediate adsorption. The CuIr electrode achieved 87% NH3 selectivity and 6.1 mmol h−1 cm−2 yield at −0.22 V vs. RHE, outperforming annealed counterparts.

4.5. Volatile Organic Compounds (VOCs) Degradation

Volatile organic compounds are toxic or carcinogenic pollutants (e.g., benzene, formaldehyde, chlorofluorocarbons) that pose significant risks to environmental and human health. These compounds are commonly emitted from industrial emissions, vehicle exhaust, and household products. Their degradation—typically transforming VOCs into benign products (e.g., CO2, H2O)—is critical for air purification. Defect engineering enhances VOC degradation efficiency by tailoring atomic-scale defects (e.g., vacancies, doping, lattice distortion) to optimize surface oxygen activity, inhibit charge recombination, and promote free radical generation.
Embedding atomically dispersed active sites within nanostructured carbon frameworks presents a promising approach for advanced VOC degradation via reactive oxygen species generation. Wei et al. designed a ZnO-carbon nanoreactor with carbon vacancy-rich Fe-N4 sites via a self-carbothermal reduction strategy [77]. Carbon vacancies near Fe-N4 sites facilitated C–O bond formation, enabled O to obtain more electron transfered from C at C–O and reduced the rate-determining step’s energy barrier, enhancing the oxygenation and reduction activity. The optimized Fe-NCv-900 activated peroxymonosulfate to generate reactive oxygen species, achieving 97% phenol removal within 5 min (13.5-fold faster than origin Fe-NC-900). The catalyst also exhibited broad pH adaptability (3–9).
Engineering high-index facets and defect-rich metal oxide supports offers a robust strategy for boosting redox activity and pollutant degradation selectivity. Sun et al. synthesized Ru-loaded high-index anatase TiO2 {201} (Ru/{201}-TiO2) rich in oxygen vacancies (VO) and Ti3+ defects [78]. The strong metal-support interaction and defect-mediated electron transfer enabled efficient O2/H2O activation and ROS generation. Ru/{201}-TiO2 achieved 100% chlorobenzene conversion without polychlorinated byproducts, with excellent stability for olefinic and alkenes CVOCs.
Amorphous metal oxide catalysts with tailored electronic structures and abundant defect sites offer promising platforms for efficient VOC degradation. Zhu et al. developed amorphous CeMnx catalysts, where CeO2 dissolution into MnO2 lattices created abundant defects [79]. Ce doping increased the conduction band’s density of states, narrowed the bandgap, and reduced the work function, enhancing electron transfer for chlorobenzene (CB) degradation. At 200 °C, CeMn0.16 achieved higher CO2 yield and Cl selectivity than MnO2.
In another study, Zhong et al. introduced pentagon defects into Co-N4-anchored carbon (C5-Co-N4) via selective pyridine N-etching [80]. The defects elevated Co’s electron density and d-band center, enabling efficient PMS activation. C5-Co-N4 exhibited a turnover frequency (TOF) of 9.37 min−1 and degraded 4-chlorophenol (4-CP) with 0.38 wt.% Co loading, outperforming conventional catalysts.

4.6. Other Catalytic Reactions

Defect engineering has emerged as a pivotal strategy to enhance diverse catalytic reactions, including selective toluene oxidation and glycerol oxidation. Shi et al. developed Ni-doped monolayered Bi2WO6 nanosheets (Ni/BWO) with varying Ni mass fractions for selective photocatalytic toluene oxidation [81]. Ni dopants mediate cascade active units at the atomic scale, comprising unsaturated metal atoms and spatially separated Bi···O frustrated Lewis pairs (FLPs). These FLPs activate toluene C–H bonds via concerted Bi···C and O···H coordination, with O2 adsorbed at unsaturated W sites. This configuration bridges charge transfer, accelerating toluene oxidation by photoinduced radicals. Additionally, FLPs and W sites establish a unique pathway for C–H dehydrogenation, O2 activation, and oxygen transfer. At 1.8 wt.% Ni, Ni/BWO achieves a toluene conversion rate of 4560 μmol g−1 h−1 with over 90% benzaldehyde selectivity.
Wu et al. engineered a spatially asymmetric defect structure in CuCo2O4 by partially refilling sulfur into high-density oxygen vacancies (HVO-S) [82]. This optimization lowers the glycerol adsorption barrier and thermodynamically favors C–C bond cleavage. The HVO-S catalyst delivers a Faradaic efficiency (FE) of 98.5% for formate at 1.36 V vs. RHE, outperforming low-vacancy catalysts (HVO: 75.3%). DFT calculations reveal that asymmetric defects enhance local charge transfer, reduce glycerol activation barriers, and facilitate secondary C–C cleavage. The strategy also improves C–C scission in ethylene glycol and glucose electrooxidation, demonstrating universal applicability. In a paired electrolysis system coupling glycerol oxidation with nitrate reduction (NO3-RR), HVO-S (anode) and CuCo3N (cathode) achieve a cell voltage of 1.27 V at 10 mA·cm−2, with 94% FE for formate, 99% FE for NH3, and 24% energy savings compared to RuO2//Pt/C.

5. Conclusions and Perspectives

Defect engineering has emerged as a transformative strategy for precise modulation of material electronic structures, offering capabilities far beyond the intrinsic limitation of conventional perfect crystals, demonstrating immense potential in catalysis. By introducing vacancies, dopants, grain boundaries, and other defect configurations, it enables atomic-scale tailoring of band structures, charge transport, surface adsorption, and spin polarization. This review systematically elucidates the coupling mechanisms between defects and electronic properties, summarizes advanced characterization techniques and computational tools, and highlights recent breakthroughs in defect-engineered catalysis. Key mechanisms include defect-induced energy levels, Fermi level shifts, coordination environment tuning, lattice distortion, charge trapping, metal-support interactions, spin-state transitions, ferromagnetic coupling, defect dynamics, and interface engineering.
Despite significant progress, critical challenges remain:
  • Precise Structure–Property Relationships: Quantitative models linking defect types (point/line/surface), spatial distribution (monodispersed/clustered), and concentrations to electronic responses are lacking. Breakthroughs require ultrafast spectroscopy (e.g., femtosecond X-ray absorption) combined with machine learning to establish a “defect configuration–electronic response” database.
  • Multi-Defect Synergy: Single defects often suffer from limited functionality or instability. Cooperative defects (e.g., oxygen vacancies with metal dopants) can induce interfacial charge redistribution and strain–field coupling. Future work should focus on in situ nanofabrication for spatially ordered assembly of multi-defect systems.
  • Dynamic Defect Quantum Control: Current studies emphasize static defects, yet dynamic evolution under operando conditions (e.g., electrochemical-bias-induced vacancy migration) may trigger electronic state reorganization or quantum phase transitions. Atomic-resolution spatiotemporal characterization (e.g., in situ TEM or time-resolved X-ray probes) is essential.
  • Sustainable Application: Advancing green synthesis and scalable manufacturing will align defect engineering with global sustainability and carbon neutrality goals.
  • AI-Driven Design: Graph neural networks could predict defect-induced electronic perturbations, while automated platforms like droplet microreactors enable tailored synthesis for energy catalysis or optoelectronics.
Defect engineering is fundamentally reshaping the paradigm of materials science, transitioning from “structure determines properties” to a forward-looking perspective of “defects create functionality”. With the integration of atomic-scale manufacturing and multiscale theory, the field will evolve from “observing defects” to “programming defects,” offering transformative solutions to global energy and environmental crises.

Author Contributions

Conceptualization, Z.Z., Y.W., T.G. and P.H.; resources, Z.Z.; writing—original draft preparation, Z.Z.; writing—review and editing, Z.Z., Y.W., T.G. and P.H.; visualization, Z.Z.; supervision, T.G. and P.H.; project administration, T.G. and P.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study.

Acknowledgments

The authors would like to thank the Beihang university for their co operation.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zheng, J.; Meng, D.; Guo, J.; Liu, X.; Zhou, L.; Wang, Z. Defect Engineering for Enhanced Electrocatalytic Oxygen Reaction on Transition Metal Oxides: The Role of Metal Defects. Adv. Mater. 2024, 36, 2405129. [Google Scholar] [CrossRef] [PubMed]
  2. Xiang, Y.; Sun, Y.; Liu, Y.; Cai, Q.; Zhao, J. Boosting the Hydrogen Evolution Reaction on NiTe Monolayers via Defect Engineering: A Computational Study. Inorg. Chem. Front. 2024, 11, 1279–1288. [Google Scholar] [CrossRef]
  3. Yang, J.; An, L.; Wang, S.; Zhang, C.; Luo, G.; Chen, Y.; Yang, H.; Wang, D. Defects Engineering of Layered Double Hydroxide-Based Electrocatalyst for Water Splitting. Chin. J. Catal. 2023, 55, 116–136. [Google Scholar] [CrossRef]
  4. Zhang, H.; Ren, Y.; Yuan, Z.; Kang, N.; Mehdi, S.; Xing, C.; Liu, X.; Fan, Y.; Li, B.; Liu, B. Interfacial Active Sites on Co-Co2C@carbon Heterostructure for Enhanced Catalytic Hydrogen Generation. Rare Met. 2023, 42, 1935–1945. [Google Scholar] [CrossRef]
  5. Lu, W.; Yan, L.; Ye, W.; Ning, J.; Zhong, Y.; Hu, Y. Defect Engineering of Electrode Materials towards Superior Reaction Kinetics for High-Performance Supercapacitors. J. Mater. Chem. A 2022, 10, 15267–15296. [Google Scholar] [CrossRef]
  6. Gao, Y.; Huang, T.; Wu, Z.; Shi, T.; Xie, W. Band Structure Engineering in 2D BA2PbI4/InSe Perovskite Heterostructures and Superlattices. Appl. Phys. Lett. 2025, 126, 032104. [Google Scholar] [CrossRef]
  7. Han, G.; Sun, M.; Zhao, R.; Junaid, Q.M.; Hou, G.; Mohmand, N.Z.K.; Jia, C.; Liu, Y.; Van Der Voort, P.; Shi, L.; et al. Defect Engineered Ti-MOFs and Their Applications. Chem. Soc. Rev. 2025, 54, 5081–5107. [Google Scholar] [CrossRef]
  8. Deng, Z.; Liu, Y.; Lin, J.; Chen, W. Rational Design and Energy Catalytic Application of High-Loading Single-Atom Catalysts. Rare Met. 2024, 43, 4844–4866. [Google Scholar] [CrossRef]
  9. Muhammad, P.; Zada, A.; Rashid, J.; Hanif, S.; Gao, Y.; Li, C.; Li, Y.; Fan, K.; Wang, Y. Defect Engineering in Nanocatalysts: From Design and Synthesis to Applications. Adv. Funct. Mater. 2024, 34, 2314686. [Google Scholar] [CrossRef]
  10. Dou, Y.; He, C.; Zhang, L.; Yin, H.; Al-Mamun, M.; Ma, J.; Zhao, H. Approaching the Activity Limit of CoSe2 for Oxygen Evolution via Fe Doping and Co Vacancy. Nat. Commun. 2020, 11, 1664. [Google Scholar] [CrossRef]
  11. Wang, X.; Li, S.; Yuan, Z.; Sun, Y.; Tang, Z.; Gao, X.; Zhang, H.; Li, J.; Wang, S.; Yang, D.; et al. Optimizing Electrocatalytic Nitrogen Reduction via Interfacial Electric Field Modulation: Elevating d-Band Center in WS2–WO3 for Enhanced Intermediate Adsorption. Angew. Chem. 2023, 135, e202303794. [Google Scholar] [CrossRef]
  12. Ren, B.; Huang, J.; Yuan, Q.; Liu, X.; Li, P.; Zhu, G.; Xu, W.; Dong, B. Tungsten Defect-Induced Surface Reconstruction of Ce2W2O9 Arrays for Enhanced Oxygen Evolution Reaction Performance. Adv. Funct. Mater. 2025, 2501328. [Google Scholar] [CrossRef]
  13. Xu, F.; He, Y.; Zhang, J.; Liang, G.; Liu, C.; Yu, J. Prolonging Charge Carrier Lifetime via Intraband Defect Levels in S-Scheme Heterojunctions for Artificial Photosynthesis. Angew. Chem. Int. Ed. 2025, 64, e202414672. [Google Scholar] [CrossRef] [PubMed]
  14. Yang, W.; Sun, K.; Wan, J.; Ma, Y.-A.; Wang, Y.; Liu, L.; Zhu, B.; Fu, F. Dual-Site Oxygen Activation for Enhanced Photocatalytic Aerobic Oxidation by S-Scheme Ni2P/Bi3O4Br-OVs Heterojunction. Chem. Eng. J. 2023, 452, 139425. [Google Scholar] [CrossRef]
  15. Ma, X.; Ding, B.; Yang, Z.; Liu, S.; Liu, Z.; Meng, Q.; Chen, H.; Li, J.; Li, Z.; Ma, P.; et al. Sulfur-Vacancy-Engineered Two-Dimensional Cu@SnS2–x Nanosheets Constructed via Heterovalent Substitution for High-Efficiency Piezocatalytic Tumor Therapy. J. Am. Chem. Soc. 2024, 146, 21496–21508. [Google Scholar] [CrossRef] [PubMed]
  16. Xing, J.; Huang, X.; Yong, X.; Li, X.; Li, J.; Wang, J.; Wang, N.; Hao, H. N-Doped Synergistic Porous Thin-Walled g-C3N4 Nanotubes for Efficient Tetracycline Photodegradation. Chem. Eng. J. 2023, 455, 140570. [Google Scholar] [CrossRef]
  17. Lin, H.; Xin, X.; Xu, L.; Li, P.; Chen, D.; Turkevych, V.; Li, Y.; Wang, H.; Xu, J.; Wang, L. Defect-Mediated Fermi Level Modulation Boosting Photo-Activity of Spatially-Ordered S-Scheme Heterojunction. J. Colloid Interface Sci. 2024, 676, 310–322. [Google Scholar] [CrossRef]
  18. Liu, C.; Mao, S.; Shi, M.; Hong, X.; Wang, D.; Wang, F.; Xia, M.; Chen, Q. Enhanced Photocatalytic Degradation Performance of BiVO4/BiOBr through Combining Fermi Level Alteration and Oxygen Defect Engineering. Chem. Eng. J. 2022, 449, 137757. [Google Scholar] [CrossRef]
  19. Yu, Z.; Lv, S.; Yao, Q.; Fang, N.; Xu, Y.; Shao, Q.; Pao, C.; Lee, J.; Li, G.; Yang, L.; et al. Low-Coordinated Pd Site within Amorphous Palladium Selenide for Active, Selective, and Stable H2O2 Electrosynthesis. Adv. Mater. 2023, 35, 2208101. [Google Scholar] [CrossRef]
  20. Zhao, W.; Cui, C.; Xu, Y.; Liu, Q.; Zhang, Y.; Zhang, Z.; Lu, S.; Rong, Z.; Li, X.; Fang, Y.; et al. Triggering Pt Active Sites in Basal Plane of Van Der Waals PtTe2 Materials by Amorphization Engineering for Hydrogen Evolution. Adv. Mater. 2023, 35, 2301593. [Google Scholar] [CrossRef]
  21. Wang, X.; Gao, R.; Fan, G.; Guo, Y.; Han, C.; Gao, Y.; Shen, A.; Wu, L.; Gu, X. Dual Defects-Induced Iron Single Atoms Immobilized in Metal-Organic Framework-Derived Hollow BiOBr Microtubes for Low-Barrier Photocatalytic Nitrogen Reduction. Angew. Chem. Int. Ed. 2025, 64, e202501297. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, H.; Song, L.; Lv, X.; Wang, H.; Zhang, F.; Hao, S.; Wei, R.; Zhang, L.; Han, Q.; Zheng, G. Low-Coordination Triangular Cu3 Motif Steers CO2 Photoreduction to Ethanol. Angew. Chem. Int. Ed. 2025, 64, e202500928. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, J.; Qian, J.; Ran, J.; Xi, P.; Yang, L.; Gao, D. Engineering Lower Coordination Atoms onto NiO/Co3O4 Heterointerfaces for Boosting Oxygen Evolution Reactions. ACS Catal. 2020, 10, 12376–12384. [Google Scholar] [CrossRef]
  24. Liu, Y.; Wang, Y.; Li, H.; Kim, M.G.; Duan, Z.; Talat, K.; Lee, J.Y.; Wu, M.; Lee, H. Effectiveness of Strain and Dopants on Breaking the Activity-Stability Trade-off of RuO2 Acidic Oxygen Evolution Electrocatalysts. Nat. Commun. 2025, 16, 1717. [Google Scholar] [CrossRef]
  25. Li, Z.; Wang, Y.; Liu, H.; Feng, Y.; Du, X.; Xie, Z.; Zhou, J.; Liu, Y.; Song, Y.; Wang, F.; et al. Electroreduction-Driven Distorted Nanotwins Activate Pure Cu for Efficient Hydrogen Evolution. Nat. Mater. 2025, 24, 424–432. [Google Scholar] [CrossRef]
  26. Zhang, Y.; Chen, M.; He, X.; Zhao, E.; Liang, H.; Shang, J.; Liu, K.; Chen, J.; Zuo, S.; Zhou, M. Intrinsic Strain of Defect Sites Steering Chlorination Reaction for Water Purification. Nat. Commun. 2025, 16, 2652. [Google Scholar] [CrossRef]
  27. Zhang, J.; Liang, Y.; Wang, Y.; Jia, Y.; He, J.; Tang, Q.; Liu, L.; Dong, J. Manganese Atom-Induced Lattice Distortion in Fe3O4 to Boost Selective Transfer Hydrogenation of Unsaturated Oxygenates. ACS Catal. 2025, 15, 7505–7515. [Google Scholar] [CrossRef]
  28. Ma, Z.; Ma, X.; Zhang, L.; Cheng, H.; Shi, F. 1T-WS2/NCN Photocatalyst with Unique Trion Behavior and Cyano-Defects: Photocatalytic Hydrogen Evolution and Mechanistic Insights. Sep. Purif. Technol. 2023, 325, 124743. [Google Scholar] [CrossRef]
  29. Zhang, J.; Yang, X.; Xu, G.; Biswal, B.K.; Balasubramanian, R. Accumulation of Long-Lived Photogenerated Holes at Indium Single-Atom Catalysts via Two Coordinate Nitrogen Vacancy Defect Engineering for Enhanced Photocatalytic Oxidation. Adv. Mater. 2024, 36, 2309205. [Google Scholar] [CrossRef]
  30. Su, H.; Yin, H.; Orbell, W.; Li, Y.; Wang, G.; Wang, Y.; Mori, K.; Chen, Z.; Li, H.; Yamashita, H.; et al. Asymmetric Triple-Atom Sites Combined with Oxygen Vacancy for Selective Photocatalytic Conversion of CO2 to Propionic Acid. Angew. Chem. Int. Ed. 2025, 64, e202425446. [Google Scholar] [CrossRef]
  31. Bai, H.; Yang, Y.; Huang, Y.; Yuan, H.; Dong, M.; Ni, C.; Liu, X. Quantum Confinement and Defect Engineering Mediated “Two Birds with One Stone” Strategy for Co/Bi Self-Regeneration and Effective Photogenerated Carrier Separation. Appl. Catal. B Environ. Energy 2025, 365, 125001. [Google Scholar] [CrossRef]
  32. Zhang, S.; Zhang, Z.; Si, Y.; Li, B.; Deng, F.; Yang, L.; Liu, X.; Dai, W.; Luo, S. Gradient Hydrogen Migration Modulated with Self-Adapting S Vacancy in Copper-Doped ZnIn2S4 Nanosheet for Photocatalytic Hydrogen Evolution. ACS Nano 2021, 15, 15238–15248. [Google Scholar] [CrossRef] [PubMed]
  33. Xie, M.; Dai, F.; Guo, H.; Du, P.; Xu, X.; Liu, J.; Zhang, Z.; Lu, X. Improving Electrocatalytic Nitrogen Reduction Selectivity and Yield by Suppressing Hydrogen Evolution Reaction via Electronic Metal-Support Interaction. Adv. Energy Mater. 2023, 13, 2203032. [Google Scholar] [CrossRef]
  34. Zhu, H.; Wang, Y.; Jiang, Z.; Deng, B.; Xin, Y.; Jiang, Z. Defect Engineering Promoted Ultrafine Ir Nanoparticle Growth and Sr Single-Atom Adsorption on TiO2 Nanowires to Achieve High-Performance Overall Water Splitting in Acidic Media. Adv. Energy Mater. 2024, 14, 2303987. [Google Scholar] [CrossRef]
  35. Ren, G.; Zhao, J.; Zhao, Z.; Li, Z.; Wang, L.; Zhang, Z.; Li, C.; Meng, X. Defects-Induced Single-Atom Anchoring on Metal-Organic Frameworks for High-Efficiency Photocatalytic Nitrogen Reduction. Angew. Chem. Int. Ed. 2024, 63, e202314408. [Google Scholar] [CrossRef] [PubMed]
  36. Zhang, Y.; Chen, F.; Yang, X.; Guo, Y.; Zhang, X.; Dong, H.; Wang, W.; Lu, F.; Lu, Z.; Liu, H.; et al. Electronic Metal-Support Interaction Modulates Cu Electronic Structures for CO2 Electroreduction to Desired Products. Nat. Commun. 2025, 16, 1956. [Google Scholar] [CrossRef]
  37. Singhvi, C.; Sharma, G.; Verma, R.; Paidi, V.K.; Glatzel, P.; Paciok, P.; Patel, V.B.; Mohan, O.; Polshettiwar, V. Tuning the Electronic Structure and SMSI by Integrating Trimetallic Sites with Defective Ceria for the CO2 Reduction Reaction. Proc. Natl. Acad. Sci. USA 2025, 122, e2411406122. [Google Scholar] [CrossRef]
  38. Wang, Y.; Li, S.; Hou, X.; Cui, T.; Zhuang, Z.; Zhao, Y.; Wang, H.; Wei, W.; Xu, M.; Fu, Q.; et al. Low-Spin Fe3+ Evoked by Multiple Defects with Optimal Intermediate Adsorption Attaining Unparalleled Performance in Water Oxidation. Adv. Mater. 2024, 36, 2412598. [Google Scholar] [CrossRef]
  39. Pei, W.; Ma, X.; Wu, Y.; Wang, Y.; Zhou, L.; Lei, J.; Yamashita, H.; Zhang, J. A-Site Defect Regulates d-Band Center in Perovskite-Type Catalysts Enhancing Photo-Assisted Peroxymonosulfate Activation for Levofloxacin Removal via High-Valent Iron-Oxo Species. Appl. Catal. B Environ. Energy 2025, 371, 125273. [Google Scholar] [CrossRef]
  40. Huang, G.; Xia, Y.; Yang, F.; Long, W.; Liu, J.; Liao, J.; Zhang, M.; Liu, J.; Lan, Y. On-Off Switching of a Photocatalytic Overall Reaction through Dynamic Spin-State Transition in a Hofmann Clathrate System. J. Am. Chem. Soc. 2023, 145, 26863–26870. [Google Scholar] [CrossRef]
  41. Sun, T.; Tang, Z.; Zang, W.; Li, Z.; Li, J.; Li, Z.; Cao, L.; Dominic Rodriguez, J.S.; Mariano, C.O.M.; Xu, H.; et al. Ferromagnetic Single-Atom Spin Catalyst for Boosting Water Splitting. Nat. Nanotechnol. 2023, 18, 763–771. [Google Scholar] [CrossRef]
  42. Wang, T.; Jiang, H.; Zhang, C.; Li, J.; Xu, R.; Pan, F.; Chen, R.; Cai, C.; Liu, S.; Zhou, Y.; et al. Vacancy Defect Activation Spin Magnetic Effect of Ni(OH)2 Enhanced Oxygen Catalysis. Int. J. Hydrogen Energy 2024, 72, 201–208. [Google Scholar] [CrossRef]
  43. Li, M.; Li, H.; Fan, H.; Liu, Q.; Yan, Z.; Wang, A.; Yang, B.; Wang, E. Engineering Interfacial Sulfur Migration in Transition-Metal Sulfide Enables Low Overpotential for Durable Hydrogen Evolution in Seawater. Nat. Commun. 2024, 15, 6154. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, L.; Hung, S.; Zhao, S.; Wang, Y.; Bi, S.; Li, S.; Ma, J.; Zhang, C.; Zhang, Y.; Li, L.; et al. Modulating the Covalency of Ru–O Bonds by Dynamic Reconstruction for Efficient Acidic Oxygen Evolution. Nat. Commun. 2025, 16, 3502. [Google Scholar] [CrossRef]
  45. Alina, A.K.; Kadyrzhanov, K.K.; Kozlovskiy, A.A.; Konuhova, M.; Popov, A.I.; Shlimas, D.D.; Borgekov, D.B. WO3/ZnWO4 Microcomposites with Potential Application in Photocatalysis. Opt. Mater. 2024, 150, 115280. [Google Scholar] [CrossRef]
  46. Lisitsyn, V.M.; Musakhanov, D.A.; Korzhneva, T.G.; Strelkova, A.V.; Lisitsyna, L.A.; Golkovsky, M.G.; Zhunusbekov, A.M.; Karipbaev, J.T.; Kozlovsky, A.L. Synthesis and Characterization of Ceramics BaxMg(2-x)F4 Activated by Tungsten. GLASS Phys. Chem. 2023, 49, 288–292. [Google Scholar] [CrossRef]
  47. Huang, Q.; Xia, G.; Huang, B.; Xie, D.; Wang, J.; Wen, D.; Lin, D.; Xu, C.; Gao, L.; Wu, Z.; et al. Activating Lattice Oxygen by a Defect-Engineered Fe2O3–CeO2 Nano-Heterojunction for Efficient Electrochemical Water Oxidation. Energy Environ. Sci. 2024, 17, 5260–5272. [Google Scholar] [CrossRef]
  48. Li, H.; Zhang, J.; Deng, X.; Wang, Y.; Meng, G.; Liu, R.; Huang, J.; Tu, M.; Xu, C.; Peng, Y.; et al. Structure and Defect Engineering Synergistically Boost High Solar-to-Chemical Conversion Efficiency of Cerium Oxide/Au Hollow Nanomushrooms for Nitrogen Photofixation. Angew. Chem. Int. Ed. 2024, 63, e202316384. [Google Scholar] [CrossRef]
  49. Zu, D.; Ying, Y.; Wei, Q.; Xiong, P.; Ahmed, M.S.; Lin, Z.; Li, M.; Li, M.; Xu, Z.; Chen, G.; et al. Oxygen Vacancies Trigger Rapid Charge Transport Channels at the Engineered Interface of S-Scheme Heterojunction for Boosting Photocatalytic Performance. Angew. Chem. Int. Ed. 2024, 63, e202405756. [Google Scholar] [CrossRef]
  50. Deng, X.; Zhang, J.; Qi, K.; Liang, G.; Xu, F.; Yu, J. Ultrafast Electron Transfer at the In2O3/Nb2O5 S-Scheme Interface for CO2 Photoreduction. Nat. Commun. 2024, 15, 4807. [Google Scholar] [CrossRef]
  51. Wu, L.; Ning, M.; Xing, X.; Wang, Y.; Zhang, F.; Gao, G.; Song, S.; Wang, D.; Yuan, C.; Yu, L.; et al. Boosting Oxygen Evolution Reaction of (Fe,Ni)OOH via Defect Engineering for Anion Exchange Membrane Water Electrolysis Under Industrial Conditions. Adv. Mater. 2023, 35, 2306097. [Google Scholar] [CrossRef] [PubMed]
  52. Ding, J.; Guo, D.; Wang, N.; Wang, H.; Yang, X.; Shen, K.; Chen, L.; Li, Y. Defect Engineered Metal-Organic Framework with Accelerated Structural Transformation for Efficient Oxygen Evolution Reaction. Angew. Chem. Int. Ed. 2023, 62, e202311909. [Google Scholar] [CrossRef]
  53. Bi, F.; Meng, Q.; Zhang, Y.; Chen, H.; Jiang, B.; Lu, H.; Liu, Q.; Zhang, H.; Wu, Z.; Weng, X. Engineering Triple O-Ti-O Vacancy Associates for Efficient Water-Activation Catalysis. Nat. Commun. 2025, 16, 851. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, J.; Wang, Z.; Lin, X.; Gao, X.; Wang, Q.; Huang, R.; Ruan, Y.; Xu, H.; Tian, L.; Ling, C.; et al. Mn-Ce Symbiosis: Nanozymes with Multiple Active Sites Facilitate Scavenging of Reactive Oxygen Species (ROS) Based on Electron Transfer and Confinement Anchoring. Angew. Chem. Int. Ed. 2025, 64, e202416686. [Google Scholar] [CrossRef] [PubMed]
  55. Wu, X.; Li, Z.; Wang, H.; Wang, J.; Xi, G.; Zhao, X.; Zhang, C.; Liao, W.; Ho, J.C. Defect-Engineered Multi-Intermetallic Heterostructures as Multisite Electrocatalysts for Efficient Water Splitting. Adv. Sci. 2025, 2502244. [Google Scholar] [CrossRef]
  56. Li, Z.; Mao, C.; Pei, Q.; Duchesne, P.N.; He, T.; Xia, M.; Wang, J.; Wang, L.; Song, R.; Ali, F.M.; et al. Engineered Disorder in CO2 Photocatalysis. Nat. Commun. 2022, 13, 7205. [Google Scholar] [CrossRef]
  57. Yang, P.; Li, J.; Vlachos, D.G.; Caratzoulas, S. Tuning Active Site Flexibility by Defect Engineering of Graphene Ribbon Edge-hosted Fe−N3 Sites. Angew. Chem. Int. Ed. 2024, 63, e202311174. [Google Scholar] [CrossRef]
  58. Liu, M.; Li, Y.; Yang, L.; Zhao, P.; Li, J.; Tian, L.; Cao, D.; Chen, Z. Defect-Triggered Orbital Hybridization in FeMn Dual-Atom Catalysts Toward Sabatier-Optimized Oxygen Reduction. Angew. Chem. Int. Ed. 2025, e202505268. [Google Scholar] [CrossRef]
  59. Zhou, J.; Qiu, S.; Hou, X.; Ni, T.; Zhang, C.; Dai, S.; Wang, X.; Wang, G.; Jiang, H.; Huang, M. Defect-Driven Stepwise Activation of Metal-Organic Frameworks Toward Industrial-Level Anion Exchange Membrane Water Electrolysis. Angew. Chem. Int. Ed. 2025, e202503787. [Google Scholar] [CrossRef]
  60. Chen, Y.; Xu, J.; Chen, Y.; Wang, L.; Jiang, S.; Xie, Z.; Zhang, T.; Munroe, P.; Peng, S. Rapid Defect Engineering in FeCoNi/FeAl2O4 Hybrid for Enhanced Oxygen Evolution Catalysis: A Pathway to High-Performance Electrocatalysts. Angew. Chem. Int. Ed. 2024, 63, e202405372. [Google Scholar] [CrossRef]
  61. Lin, J.; Ban, T.; Li, T.; Sun, Y.; Zhou, S.; Li, R.; Su, Y.; Kasemchainan, J.; Gao, H.; Shi, L.; et al. Machine-learning-assisted Intelligent Synthesis of UiO-66(Ce): Balancing the Trade-off between Structural Defects and Thermal Stability for Efficient Hydrogenation of Dicyclopentadiene. Mater. Genome Eng. Adv. 2024, 2, e61. [Google Scholar] [CrossRef]
  62. Pei, X.; Bian, J.; Zhang, W.; Hu, Z.; Ng, Y.; Dong, Y.; Zhai, X.; Wei, Z.; Liu, Y.; Deng, J.; et al. Overcoming Defect Limitations in Photocatalysis: Boron-Incorporation Engineered Crystalline Red Phosphorus for Enhanced Hydrogen Production. Adv. Funct. Mater. 2024, 34, 2400542. [Google Scholar] [CrossRef]
  63. Liu, H.; Zhang, S.; Chai, Y.; Dong, B. Ligand Modulation of Active Sites to Promote Cobalt-Doped 1T-MoS2 Electrocatalytic Hydrogen Evolution in Alkaline Media. Angew. Chem. Int. Ed. 2023, 62, e202313845. [Google Scholar] [CrossRef]
  64. Zhang, Y.; Yun, S.; Dang, J.; Dang, C.; Yang, G.; Wang, Y.; Liu, Z.; Deng, Y. Defect Engineering via Ternary Nonmetal Doping Boosts the Catalytic Activity of ZIF-Derived Carbon-Based Metal-Free Catalysts for Photovoltaics and Water Splitting. Mater. Today Phys. 2022, 27, 100785. [Google Scholar] [CrossRef]
  65. Zhang, Y.; Liu, M.; Zhang, L.; Lu, N.; Wang, X.; Li, Z.; Zhang, X.; Li, N.; Bu, X. Strategic Defect Engineering Enabled Efficient Oxygen Evolution Reaction in Reconstructed Metal-Organic Frameworks. Adv. Funct. Mater. 2025, 35, 2412406. [Google Scholar] [CrossRef]
  66. Zhao, X.; Wang, S.; Cao, Y.; Li, Y.; Portniagin, A.S.; Tang, B.; Liu, Q.; Kasák, P.; Zhao, T.; Zheng, X.; et al. High-Density Atomic Level Defect Engineering of 2D Fe-Based Metal-Organic Frameworks Boosts Oxygen and Hydrogen Evolution Reactions. Adv. Sci. 2024, 11, 2405936. [Google Scholar] [CrossRef]
  67. Gu, M.; Jiang, L.; Wang, H.; Chen, Q.; Hao, Y.; Li, L.; Hu, F.; Zhang, X.; Wu, Y.; Wang, G.; et al. Iodine-Induced Redirection of Active Sources in Cu-Based Catalysts during Efficient and Stable Water Oxidation. J. Am. Chem. Soc. 2025, 147, 19675–19686. [Google Scholar] [CrossRef]
  68. Zhang, R.; Pan, L.; Guo, B.; Huang, Z.; Chen, Z.; Wang, L.; Zhang, X.; Guo, Z.; Xu, W.; Loh, K.P.; et al. Tracking the Role of Defect Types in Co3O4 Structural Evolution and Active Motifs during Oxygen Evolution Reaction. J. Am. Chem. Soc. 2023, 145, 2271–2281. [Google Scholar] [CrossRef]
  69. Lai, K.; Sun, Y.; Li, N.; Gao, Y.; Li, H.; Ge, L.; Ma, T. Photocatalytic CO2-to-CH4 Conversion with Ultrahigh Selectivity of 95.93% on S-Vacancy Modulated Spatial In2S3/In2O3 Heterojunction. Adv. Funct. Mater. 2024, 34, 2409031. [Google Scholar] [CrossRef]
  70. Wang, Y.; Hu, J.; Ge, T.; Chen, F.; Lu, Y.; Chen, R.; Zhang, H.; Ye, B.; Wang, S.; Zhang, Y.; et al. Gradient Cationic Vacancies Enabling Inner-To-Outer Tandem Homojunctions: Strong Local Internal Electric Field and Reformed Basic Sites Boosting CO2 Photoreduction. Adv. Mater. 2023, 35, 2302538. [Google Scholar] [CrossRef]
  71. Wang, Z.; Cai, R.; Zhang, Y.; Huang, X.; Bi, Y. Dynamic Crystal-Structure and Active-Site of Defective ZnAl-Catalysts During CO2 Photoreduction. Angew. Chem. 2025, e202502834. [Google Scholar] [CrossRef]
  72. A, B.; Jin, X.; Wang, M.; Wang, Y.; Chen, W.; Wei, Z.; Du, Z.; Liu, X.; Wang, Y.; Zhang, L. Steering CO2 Electroreduction toward C2+ Products by Defect-Engineered Coordination Microenvironments of Cu-MOFs. Chem. Eng. J. 2024, 500, 157076. [Google Scholar] [CrossRef]
  73. Ding, Y.; Zhang, S.; Liu, Y.; Liu, Y.; Zheng, H.; Qing, L.; Song, Y.; Ma, Z.; Zhang, L.; Liu, T. Defect-Enriched Cobalt-Based Coordination Polymers for Selective and Efficient Nitrate Electroreduction to Ammonia. Adv. Funct. Mater. 2025, 2422339. [Google Scholar] [CrossRef]
  74. Yuan, J.; Feng, W.; Zhang, Y.; Xiao, J.; Zhang, X.; Wu, Y.; Ni, W.; Huang, H.; Dai, W. Unraveling Synergistic Effect of Defects and Piezoelectric Field in Breakthrough Piezo-Photocatalytic N2 Reduction. Adv. Mater. 2024, 36, 2303845. [Google Scholar] [CrossRef]
  75. Sun, Y.; Lu, G.; Wang, Z.; Li, X.; Li, Y.; Sui, N.L.D.; Fan, W.; Wang, A.; Yuan, B.; Wang, J.; et al. Size Effect of Surface Defects Dictates Reactivity for Nitrogen Electrofixation. Angew. Chem. Int. Ed. 2025, 64, e202425112. [Google Scholar] [CrossRef]
  76. Guo, X.; Yu, J.; Ren, S.; Gao, R.; Wu, L.; Wang, L. Controlled Defective Engineering on CuIr Catalyst Promotes Nitrate Selective Reduction to Ammonia. ACS Nano 2024, 18, 24252–24261. [Google Scholar] [CrossRef] [PubMed]
  77. Wei, S.; Sun, Y.; Qiu, Y.; Li, A.; Chiang, C.; Xiao, H.; Qian, J.; Li, Y. Self-Carbon-Thermal-Reduction Strategy for Boosting the Fenton-like Activity of Single Fe-N4 Sites by Carbon-Defect Engineering. Nat. Commun. 2023, 14, 7549. [Google Scholar] [CrossRef]
  78. Sun, B.; Li, Q.; Su, G.; Song, M.; Ma, C.; Pang, J.; Zhao, X.; Meng, J.; Shi, B. Sustainable and Efficient Catalytic Oxidation of Chlorinated Volatile Organic Compounds over Ru-Loaded Facet-Engineered {201}-TiO2 Catalyst with Tuned Defects. Appl. Catal. B Environ. Energy 2025, 361, 124582. [Google Scholar] [CrossRef]
  79. Zhu, Y.; Liang, W.; Zhang, C.; Bin, F.; Tao, Q. Defect-Rich Regulatory Activity Strategy: Disordered Structure for Enhanced Catalytic Interfacial Reaction of Chlorobenzene. Environ. Sci. Technol. 2024, 58, 19385–19396. [Google Scholar] [CrossRef]
  80. Zhong, H.; Wang, J.; Zhang, B.; Liu, Y.; Gong, Z.; Xiao, X.; Yu, J.; Hou, Y.; Tao, Y.; Yang, H.; et al. Selective Introduction of Pentagon Defects into Co–N4 Sites for Boosting Fenton-Like Activity. Adv. Funct. Mater. 2025, 2501208. [Google Scholar] [CrossRef]
  81. Shi, Y.; Li, P.; Chen, H.; Wang, Z.; Song, Y.; Tang, Y.; Lin, S.; Yu, Z.; Wu, L.; Yu, J.C.; et al. Photocatalytic Toluene Oxidation with Nickel-Mediated Cascaded Active Units over Ni/Bi2WO6 Monolayers. Nat. Commun. 2024, 15, 4641. [Google Scholar] [CrossRef] [PubMed]
  82. Wu, L.; Wu, Q.; Han, Y.; Zhang, D.; Zhang, R.; Song, N.; Fang, Y.; Liu, H.; Wang, M.; Chen, J.; et al. Constructing Asymmetric Defects Pairs in Electrocatalysts for Efficient Glycerol Oxidation. J. Am. Chem. Soc. 2025, 147, 18033–18043. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Total density of state (TDOS) and projected density of state (PDOS) of WO3 with various concentrations of OVs [13]. Reprinted with permission from Ref. [13]. © 2024 Wiley-VCH GmbH.
Figure 1. Total density of state (TDOS) and projected density of state (PDOS) of WO3 with various concentrations of OVs [13]. Reprinted with permission from Ref. [13]. © 2024 Wiley-VCH GmbH.
Catalysts 15 00651 g001
Figure 2. (a,b) Calculated work functions of SnS2 and Cu@SnS2−x. The work function is the difference between the energy of the fixed electrons in the vacuum close to the surface and the Fermi energy [15]. Reprinted with permission from Ref. [15]. © 2024 American Chemical Society.
Figure 2. (a,b) Calculated work functions of SnS2 and Cu@SnS2−x. The work function is the difference between the energy of the fixed electrons in the vacuum close to the surface and the Fermi energy [15]. Reprinted with permission from Ref. [15]. © 2024 American Chemical Society.
Catalysts 15 00651 g002
Figure 3. UPS spectra of MoSe2 prepared with varying reaction temperatures. The Φ of MoSe2 with differing synthesis temperatures can be derived as 4.53, 4.61, 4.58, and 4.56 eV for MoSe2-160, MoSe2-180, MoSe2-200, and MoSe2-220, corresponding to the EF of −4.53, −4.61, −4.58, and −4.56 eV, i.e., 0.09, 0.17, 0.14, and 0.12 V vs. RHE, respectively [17]. Reprinted with permission from Ref. [17]. © 2024 Elsevier Inc.
Figure 3. UPS spectra of MoSe2 prepared with varying reaction temperatures. The Φ of MoSe2 with differing synthesis temperatures can be derived as 4.53, 4.61, 4.58, and 4.56 eV for MoSe2-160, MoSe2-180, MoSe2-200, and MoSe2-220, corresponding to the EF of −4.53, −4.61, −4.58, and −4.56 eV, i.e., 0.09, 0.17, 0.14, and 0.12 V vs. RHE, respectively [17]. Reprinted with permission from Ref. [17]. © 2024 Elsevier Inc.
Catalysts 15 00651 g003
Figure 4. The possible mechanism for OTC degradation in the 20% BVB and 20% BVB-Ov system [18]. Reprinted with permission from Ref. [18]. © 2022 Elsevier B.V.
Figure 4. The possible mechanism for OTC degradation in the 20% BVB and 20% BVB-Ov system [18]. Reprinted with permission from Ref. [18]. © 2022 Elsevier B.V.
Catalysts 15 00651 g004
Figure 5. Simulated local structure of a-PdSe2 NPs and c-PdSe2 NPs. Note: CN is the coordination number. Blue, Pd; Pink, Se; Red, O; White, H [19]. Reprinted with permission from Ref. [19]. © 2022 Wiley-VCH GmbH.
Figure 5. Simulated local structure of a-PdSe2 NPs and c-PdSe2 NPs. Note: CN is the coordination number. Blue, Pd; Pink, Se; Red, O; White, H [19]. Reprinted with permission from Ref. [19]. © 2022 Wiley-VCH GmbH.
Catalysts 15 00651 g005
Figure 6. (a,b) FT-EXAFS spectra and corresponding fitting curves of FeSA-Bi/h-BiOBr-VO, Br and FeSA-Bi/BiOBr-VBr, respectively. The illustrations show the fitting results for the first shell manifests [21]. Reprinted with permission from Ref. [21]. © 2025 Wiley-VCH GmbH.
Figure 6. (a,b) FT-EXAFS spectra and corresponding fitting curves of FeSA-Bi/h-BiOBr-VO, Br and FeSA-Bi/BiOBr-VBr, respectively. The illustrations show the fitting results for the first shell manifests [21]. Reprinted with permission from Ref. [21]. © 2025 Wiley-VCH GmbH.
Catalysts 15 00651 g006
Figure 7. The crystal RuO2 model (top) is endowed by tensile strain with elongated Ru–O bonding (TS-RuO2, left), and the extrinsic Sr and Ta doping (SrTaRuO2, right) modulates electronic redistribution. Combining two strategies obtained TS-SrTaRuO2 (bottom) [24]. Reprinted with permission from Ref. [24]. © 2025 Yang Liu et al.
Figure 7. The crystal RuO2 model (top) is endowed by tensile strain with elongated Ru–O bonding (TS-RuO2, left), and the extrinsic Sr and Ta doping (SrTaRuO2, right) modulates electronic redistribution. Combining two strategies obtained TS-SrTaRuO2 (bottom) [24]. Reprinted with permission from Ref. [24]. © 2025 Yang Liu et al.
Catalysts 15 00651 g007
Figure 8. (a) Schematic diagram of the transient absorption (TA) signal probed at 465 nm. The fast decay component (τ1) corresponds to the trapping of electrons from the LUMO into near band-edge trap states of CN. The slower decay component (τ2) reflects the recombination between the trap band electrons and holes residing in the HOMO [29]. (b) Corresponding kinetic fitting results at 480 nm over CWO and CWO-OVs. In CWO-OVs, the OV-mediated trap states can transiently capture the surrounding photogenerated electrons, resulting in an extended component (τ3 = 635.62 ps) and an increased average carrier lifetime (τave = 545.99 ps) [30]. Reprinted with permission from Ref. [29]. © 2024 Rajasekhar Balasubramanian, Basanta Kumar Biswal, Guofang Xu et al. Advanced Materials published by Wiley-VCH GmbH. Reprinted with permission from Ref. [30]. © 2025 Wiley-VCH GmbH.
Figure 8. (a) Schematic diagram of the transient absorption (TA) signal probed at 465 nm. The fast decay component (τ1) corresponds to the trapping of electrons from the LUMO into near band-edge trap states of CN. The slower decay component (τ2) reflects the recombination between the trap band electrons and holes residing in the HOMO [29]. (b) Corresponding kinetic fitting results at 480 nm over CWO and CWO-OVs. In CWO-OVs, the OV-mediated trap states can transiently capture the surrounding photogenerated electrons, resulting in an extended component (τ3 = 635.62 ps) and an increased average carrier lifetime (τave = 545.99 ps) [30]. Reprinted with permission from Ref. [29]. © 2024 Rajasekhar Balasubramanian, Basanta Kumar Biswal, Guofang Xu et al. Advanced Materials published by Wiley-VCH GmbH. Reprinted with permission from Ref. [30]. © 2025 Wiley-VCH GmbH.
Catalysts 15 00651 g008
Figure 9. (a) Top and side views of Ir@Sr-d-TiO2(004). (b) Schematic illustration of the OER processes on Ir@Sr-d-TiO2(004). “*” is a site on the surface. [34]. Reprinted with permission from Ref. [34]. © 2024 Wiley-VCH GmbH.
Figure 9. (a) Top and side views of Ir@Sr-d-TiO2(004). (b) Schematic illustration of the OER processes on Ir@Sr-d-TiO2(004). “*” is a site on the surface. [34]. Reprinted with permission from Ref. [34]. © 2024 Wiley-VCH GmbH.
Catalysts 15 00651 g009
Figure 10. The illustration of NiFe-LS with optimized spin state [38]. Reprinted with permission from Ref. [38]. © 2024 Wiley-VCH GmbH.
Figure 10. The illustration of NiFe-LS with optimized spin state [38]. Reprinted with permission from Ref. [38]. © 2024 Wiley-VCH GmbH.
Catalysts 15 00651 g010
Figure 11. Schematic diagram of WS2–WO3 synthesis and reaction on the surface [11]. Reprinted with permission from Ref. [11]. © 2023 Wiley-VCH GmbH.
Figure 11. Schematic diagram of WS2–WO3 synthesis and reaction on the surface [11]. Reprinted with permission from Ref. [11]. © 2023 Wiley-VCH GmbH.
Catalysts 15 00651 g011
Figure 12. EPR spectra for d-(Fe,Ni)OOH and (Fe,Ni)OOH [51]. Reprinted with permission from Ref. [51]. © 2023 Wiley-VCH GmbH.
Figure 12. EPR spectra for d-(Fe,Ni)OOH and (Fe,Ni)OOH [51]. Reprinted with permission from Ref. [51]. © 2023 Wiley-VCH GmbH.
Catalysts 15 00651 g012
Figure 13. PALS spectra for CeO2, Mn3O4, and CeOx/Mn3O4 [54]. Reprinted with permission from Ref. [54]. © 2024 Wiley-VCH GmbH.
Figure 13. PALS spectra for CeO2, Mn3O4, and CeOx/Mn3O4 [54]. Reprinted with permission from Ref. [54]. © 2024 Wiley-VCH GmbH.
Catalysts 15 00651 g013
Figure 14. XPS spectra of (a) Ni 2p, (b) Mo 3d, (c) Cu 2p, and (d) Co 2p of CoCuMoNi. The red lines are data fitting lines [55]. Reprinted with permission from Ref. [55]. © 2025 Johnny C. Ho, Wu-Gang Liao, Chen-Xu Zhang et al. Advanced Science published by Wiley-VCH GmbH.
Figure 14. XPS spectra of (a) Ni 2p, (b) Mo 3d, (c) Cu 2p, and (d) Co 2p of CoCuMoNi. The red lines are data fitting lines [55]. Reprinted with permission from Ref. [55]. © 2025 Johnny C. Ho, Wu-Gang Liao, Chen-Xu Zhang et al. Advanced Science published by Wiley-VCH GmbH.
Catalysts 15 00651 g014
Figure 15. UPS spectra of c-TiO2 (a) and a-TiO2−x(OH)y (b). The yellow lines are the tangent lines of the data lines. (c) Energy band alignment for the core@shell c-TiO2@a-TiO2−x(OH)y with respect to the absolute vacuum energy scale (AVS) [56]. Reprinted with permission from Ref. [56]. © 2022 Zhao Li et al.
Figure 15. UPS spectra of c-TiO2 (a) and a-TiO2−x(OH)y (b). The yellow lines are the tangent lines of the data lines. (c) Energy band alignment for the core@shell c-TiO2@a-TiO2−x(OH)y with respect to the absolute vacuum energy scale (AVS) [56]. Reprinted with permission from Ref. [56]. © 2022 Zhao Li et al.
Catalysts 15 00651 g015
Figure 16. Simulated Fe K-edge XANES spectra of edge-hosted Fe–N3 sites [57]. Reprinted with permission from Ref. [57]. © 2023 Wiley-VCH GmbH.
Figure 16. Simulated Fe K-edge XANES spectra of edge-hosted Fe–N3 sites [57]. Reprinted with permission from Ref. [57]. © 2023 Wiley-VCH GmbH.
Catalysts 15 00651 g016
Figure 17. k3-weighted FT-EXAFS spectra of the CoFc-MOF, D/CoFc-MOF, and the related references at the Co K–edge [59]. Reprinted with permission from Ref. [59]. © 2025 Wiley-VCH GmbH.
Figure 17. k3-weighted FT-EXAFS spectra of the CoFc-MOF, D/CoFc-MOF, and the related references at the Co K–edge [59]. Reprinted with permission from Ref. [59]. © 2025 Wiley-VCH GmbH.
Catalysts 15 00651 g017
Figure 18. (a) Aberration-corrected HAADF-STEM image obtained from boundary location of cerium oxide-AD/Au. The region 1 and 2 indicate next to and far away from the Au HNM, respectively. Corresponding EELS valence mapping showing distribution of CeIII(green) and CeIV(red). EELS spectra of (b) Ce-M and (c) O-K edges taken from region 1 (green lines) and 2 (red lines) marked in (a) [48]. Reprinted with permission from Ref. [48]. © 2023 Wiley-VCH GmbH.
Figure 18. (a) Aberration-corrected HAADF-STEM image obtained from boundary location of cerium oxide-AD/Au. The region 1 and 2 indicate next to and far away from the Au HNM, respectively. Corresponding EELS valence mapping showing distribution of CeIII(green) and CeIV(red). EELS spectra of (b) Ce-M and (c) O-K edges taken from region 1 (green lines) and 2 (red lines) marked in (a) [48]. Reprinted with permission from Ref. [48]. © 2023 Wiley-VCH GmbH.
Catalysts 15 00651 g018
Table 1. The main types of defects and their effects on electronic structures.
Table 1. The main types of defects and their effects on electronic structures.
The Main TypesThe Effects on Electronic Structures
Vacancy defectIntroduces local electronic states, forms defect energy levels in the band gap, creates charge traps, changes the energy gap, and alters the electron spin
Doping defectInduces charge transfer and orbital hybridization, regulates the position of Fermi energy levels, introduces impurity energy levels, and changes the band gap
Coordination defectInduces the rearrangement of orbital electrons and regulates the position of the Fermi energy level
Lattice distortion defectChanges the bond length, adjusts the electronic state density, and increases the center of the D-band
Interface defectForms the space charge region at the grain boundaries, generates an interfacial electric field that affects carrier migration, and alters the energy band structure
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, Z.; Wang, Y.; Guo, T.; Hu, P. The Influence of Defect Engineering on the Electronic Structure of Active Centers on the Catalyst Surface. Catalysts 2025, 15, 651. https://doi.org/10.3390/catal15070651

AMA Style

Zhang Z, Wang Y, Guo T, Hu P. The Influence of Defect Engineering on the Electronic Structure of Active Centers on the Catalyst Surface. Catalysts. 2025; 15(7):651. https://doi.org/10.3390/catal15070651

Chicago/Turabian Style

Zhang, Zhekun, Yankun Wang, Tianqi Guo, and Pengfei Hu. 2025. "The Influence of Defect Engineering on the Electronic Structure of Active Centers on the Catalyst Surface" Catalysts 15, no. 7: 651. https://doi.org/10.3390/catal15070651

APA Style

Zhang, Z., Wang, Y., Guo, T., & Hu, P. (2025). The Influence of Defect Engineering on the Electronic Structure of Active Centers on the Catalyst Surface. Catalysts, 15(7), 651. https://doi.org/10.3390/catal15070651

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop