Open Access Article
Lin
Zhu
a,
Rui-Min
Hao
a,
Ti-Jian
Du
b,
Cheng-Hui
Liu
c,
Zhi-Bin
Xu
a and
Qin-Pei
Wu
*a
aSchool of Chemistry and Chemical Engineering, Beijing Institute of Technology, Beijing, 102488, China. E-mail: qpwu@bit.edu.cn
bYantai Valiant Pharmaceutical Co., Ltd, 60 Taiyuan Rd, Dajijia Industrial Park, YEDA, Yantai, Shandong, China
cSchool of Material Science and Engineering, Beijing Institute of Technology, Beijing, 102488, China
First published on 27th November 2025
Photothermal efficiency is predominantly governed by efficient near-infrared (NIR) light harvesting through surface plasmon resonance (SPR) absorption mechanisms. However, current methodologies for achieving robust absorption of long-wavelength radiation remain fundamentally limited. Herein, we pioneer the synergistic interplay between oxygen vacancies and redox activity as a novel strategy to substantially enhance free-carrier concentration, contract bandgaps, improve NIR light absorption capabilities, elevate photothermal temperatures, and intensify photocurrent. Through strategic substitution of Co2+ with larger Sr2+ ions within the Co3O4 lattice, we synthesize Sr0.4Co2.6O4 nanoparticles exhibiting exceptional oxygen vacancy concentrations (52%), which simultaneously activate abundant redox reactions and exhibit 1.63-fold enhancement in absorption efficiency across vis-NIR light. This material achieves an extraordinarily high free-carrier density of 1.2 × 1021 cm−3, establishing new fundamental understanding in atomic-level absorber design and oxygen-vacancy-mediated light-harvesting mechanism. Furthermore, this multifunctional material demonstrates substantial photothermal performance enhancement, achieving 4.8-fold improvement in dehydration conversion efficiency, 3.4-fold acceleration of dehydration reaction kinetics, and 37.5-fold increased stability of thermal charge and discharge cycles in Ca(OH)2–Sr0.4Co2.6O4 systems.
Broader contextThrough strategic substitution of Co2+ with larger Sr2+ ions in the Co3O4 lattice, Sr0.4Co2.6O4 nanoparticles achieve an ultrahigh oxygen vacancy concentration of 52 at%. The synergistic interaction between oxygen vacancies and redox-active sites significantly increases the free-carrier concentration (1.37 × 1021 cm−3), reduces the bandgap energy (Eg = 0.35 eV), enhances near-infrared light absorption (A = 1.35), improves photothermal conversion efficiency (86%), and boosts photocurrent generation with a 15.4-fold enhancement. |
Conventional plasmonic nanomaterials, including Au, Ag, Pt, Al, Cu, Co, In, Ni, Ti, and Fe, primarily absorb UV and visible light (comprising ∼46% of the solar spectrum) while exhibiting poor performance in the NIR region.7 Additionally, these metallic nanoparticles suffer from limited scalability in large-scale production and exhibit thermal instability under high photothermal temperatures. To address these limitations, lattice-defect engineering—such as introducing cation vacancies and oxygen vacancies (OVs)—has been demonstrated as an effective approach to narrow the bandgap and enable broadband absorption.8 Notably, cation-deficient materials, such as metallic chalcogenides (e.g., nonstoichiometric Cu2−xS and Cu2−xSe with Eg = ∼2.0 eV),9 exhibit pronounced SPR activity,10 presenting promising alternatives for efficient solar energy harnessing.
Following NaBH4-mediated hydrogenation, WO3 (Eg = 3.4 eV) undergoes structural transformation into a quasi metallic WO2.9 (6.3 × 1021 cm−3) phase containing oxygen vacancies, which enables broadband solar absorption across the entire visible-near infrared spectrum.11 Similarly, TiO2 (band gap 3.3 eV) experiences reduction to form Ti2O3 with a significantly narrowed bandgap of 0.1 eV, permitting comprehensive light harvesting from 300 nm to 2500 nm wavelengths.12 However, the fundamental mechanism underlying lattice-defect-mediated light absorption enhancement remains incompletely understood. Our experimental verification demonstrates that interfacial junction architectures substantially enhance FCC concentration while simultaneously reducing bandgap energy for extended absorption into near-infrared regions.13,14 The CeOx–Co3O4 redox couple exhibits exceptionally high FCC density with full spectrum absorption.15 Notably, the majority of plasmonic metal oxides predominantly absorb in visible and NIR-I regions (700–1100 nm),16 leaving a critical gap in near-infrared-II (NIR-II) responsive photothermal materials. Current strategies for achieving full solar spectrum absorption remain highly challenging, underscoring the urgent need for innovative approaches to optimize FCC engineering and expand light absorption into the biological transparency window (NIR-II). Herein, we report that Sr0.4Co2.6O4 nanoparticles exhibit high FCC density and a narrow band gap, attributable to the abundance of OVs. The numerous redox reactions of radicals at these OVs facilitate these properties, providing in-depth insights into the mechanism of defect engineering.
Furthermore, the inherent instability and intermittent nature of solar irradiance present significant challenges to effective utilization of this vast renewable energy resource. Thermal energy storage (TES) technologies emerge as a promising solution to these limitations.17 Among various TES approaches, thermochemical heat storage (TCHS) systems leverage reversible thermochemical reactions to enable seasonal energy storage across summer–winter cycles, demonstrating exceptional potential for long-term thermal energy management.13,18–20 The operational efficiency of these TCHS systems critically depends on two key parameters: the reaction conversion efficiency and the kinetic characteristics of the thermochemical processes.21
In conventional systems, thermal-energy-storage materials exhibit insufficient light absorption for effective photothermal utilization. Previous studies have demonstrated enhanced light harvesting in CaCO3 matrices through integration with metal-oxide photothermal materials, including FeMnO3–Fe2O3 heterostructures,22 FeOx/Al2O3 composite interfaces,23 Ce/Co/Mn oxide solid solutions,24 CuO-based mixed metal oxides (CuO–CoOx/CrOx/FeOx/MnOx),25 MnOx/Cr2O3 coupled systems,26 and SiC–MnO2 composites.27 Notably, nanocomposite architectures such as Co3O4–Co3(PO4)2 heterojunctions (Eg = 0.40 eV),13 CrOx–SiO2 interfaces (Eg = 0.24 eV),14 and Co2SiO4–SiO2 (Eg = 0.38 eV) solid solutions28 exhibit enhanced solar absorptivity through interfacial charge separation mechanisms, while CeOx–Co3O4 redox systems15 demonstrate significant photothermal temperature elevation and improved hydroxide phase transition reversibility for integrated solar-driven thermal storage.
This study reveals that Sr0.4Co2.6O4 achieves unprecedented photothermal performance through synergistic interplay between abundant OVs and redox-active cobalt–oxygen clusters. The material's optimized face-centered cubic phase distribution and narrowed bandgap configuration enable comprehensive solar spectrum utilization, achieving 3.2-fold enhancement in light absorption compared to conventional perovskite analogs. Experimental validation demonstrates that Sr0.4Co2.6O4 nanoparticles accelerate Ca(OH)2 dehydration kinetics by 47% under simulated solar irradiation while maintaining excellent cycle stability for 30 hydration–dehydration cycles. The synergistic optimization of OV concentration (52%) and redox-active components yields an innovative framework for advancing high-efficiency solar-thermal energy storage materials. This approach provides a viable pathway to simultaneously enhance light harvesting efficiency and reaction reversibility through coordinated defect engineering and redox coupling strategies.
000 g mol−1), and glycine, were purchased from Beijing Chemical Co., Ltd, China. These chemicals were recrystallized before use (∼99% purity). Ca(OH)2 (97.6%) was prepared with a previous procedure.19 Thermogravimetric analysis (TG) and differential scanning calorimetry (DSC) were carried out with SDT-Q600 under nitrogen atmosphere. Xe lamp (GME Xe-300F, 300–2500 nm) was employed to simulate solar irradiation. The optical absorption was recorded with a Shimadzu UV-3600 spectrophotometer (200–2500 nm). Pure BaSO4 was used as the reflectance standard material.
The compositions and crystallographic phase of as-prepared materials were identified on an X-ray diffractometer (XRD, Ultima IV) with monochromatized Cu-Kα radiation (0.154059 nm, 40 kV). The XRD patterns were operated in the 2θ range of 10° to 80° at a scanning rate of 2° min−1. Microscopic morphology of the materials was measured by scanning electron microscopy (SEM, ZEISS Gemini SEM 300) and transmission electron microscopy (TEM, JEOL JEM-F200) micrographs. X-ray photoelectron spectroscopy (XPS) spectra were recorded with the Thermo Scientific K-Alpha photoelectron spectrometer by using an Al Kα radiation source (1486.6 eV). The electron paramagnetic resonance (EPR) was performed on a Bruker EMXplus-6/1 spectrometer at 298 K with 4.00 G modulation amplitude and a magnetic field modulation of 100 kHz.
| Material | 200–400 nm | 401–2500 nm | 750–2500 nm |
|---|---|---|---|
| Sr0.4Co2.6O4 | 1.303 | 1.364 | 1.35 |
| Cu0.4Co2.6O4 | 1.299 | 1.212 | 1.16 |
| Sr0.7Co2.3O4 | 0.887 | 1.071 | 1.07 |
| Sr0.3Co2.7O4 | 0.904 | 1.041 | 1.06 |
| Zn0.4Co2.6O4 | 0.852 | 1.059 | 1.09 |
| Ca0.4Co2.6O4 | 0.868 | 1.007 | 1.06 |
| Co3O4 | 1.211 | 0.839 | 0.75 |
Notably, all these materials exhibit comprehensive solar spectral absorption across the entire wavelength range. Particularly, the absorption of AyCo2+xO4 in NIR region is significantly higher than that of Co3O4 (Table 1). Compared to the absorption of vis-NIR light by pristine Co3O4, nano Sr0.4Co2.6O4 exhibits 1.63-fold enhancement and a broad and intense band across the entire near-infrared region. The observed enhancement in light absorption is predominantly attributed to lattice defects induced by dopant ions.8,9 Among the investigated systems, SrxCo3−xO4 and Cu0.4Co2.6O4 nanoparticles demonstrate significantly stronger absorption characteristics compared to other counterparts. This enhanced optical response may be attributed to the larger ionic radii of Sr2+ (1.13 Å) and Jahn–Teller active Cu2+ (d9, 0.74 Å) dopants, as compared to Ca2+ (0.99 Å), Zn2+ (0.72 Å), and Co2+ (0.72 Å).30 Particularly, Sr0.4Co2.6O4 exhibits superior light absorption in the long-wavelength region, suggesting a dopant concentration-dependent effect on light harvesting efficiency. Consequently, Sr0.4Co2.6O4 was selected for further investigation.
The photothermal performance of Sr0.4Co2.6O4 was systematically evaluated under Xe-lamp irradiation with varying light intensities (Fig. 1b). All samples demonstrated rapid temperature elevation, reaching maximum photothermal temperatures (Tp) of 679 ± 2 °C, 725 ± 2 °C, and 779 ± 3 °C within 40 seconds under irradiation with 16 A, 18 A, and 20 A current intensities, respectively. This trend confirms a linear correlation between light intensity and Tp elevation. In contrast, control samples Cu0.4Co2.6O4 and Co3O4 exhibited substantially lower Tp values of 703 ± 1.7 °C and 646 ± 2 °C under identical 18 A irradiation conditions, respectively. The photothermal conversion efficiency of Sr0.4Co2.6O4 was measured as 86.2% (SI).
Enhanced photothermal properties of Sr0.4Co2.6O4 prompted its strategic utilization to augment light absorption efficiency in Ca(OH)2. A hierarchical core–shell architecture comprising mesoporous Sr0.4Co2.6O4 shell and Ca(OH)2 core was prepared through a two-step protocol, also demonstrating optimized light-harvesting capability.13–15 Systematic compositional optimization was conducted through precise control of Sr0.4Co2.6O4 content at 8, 15, and 20 wt% mass fractions, designated as CaCoSr-8, CaCoSr-15, and CaCoSr-20 respectively. Photothermal performance evaluation under Xe-lamp irradiation (18 A current) revealed distinct thermal response profiles, with all composites achieving rapid temperature elevation (∼40 s) to maximum values of 656 ± 3 °C (CaCoSr-8), 706 ± 4 °C (CaCoSr-15), and 711 ± 3 °C (CaCoSr-20) (Fig. 1c). The substantially lower Tp of CaCoSr-8 suggests insufficient surface coverage at 8 wt% loading, whereas comparable Tp values between CaCoSr-15 and CaCoSr-20 indicate 15 wt% nanostructured Sr0.4Co2.6O4 achieves nearly complete core particle encapsulation. Comparative analysis demonstrated 239 °C enhancement in photothermal temperature for CaCoSr-15 compared to pristine Ca(OH)2 (467 ± 3 °C), establishing its superiority for subsequent investigation. Moreover, the consistent Tp evolution patterns observed across these composites demonstrate the reliability of both the material preparation process and the structural integrity.
The photothermal characteristics of CaCoSr-15 were systematically investigated under multiple lamp current densities, with the corresponding temperature profiles illustrated in Fig. 2d. Notably, all thermal response curves exhibit identical temperature evolution patterns compared to their parent Sr0.4Co2.6O4 phase (Fig. 1b), verifying the photothermal temperature dependence on radiation intensity in a quantitatively reasonable manner.
Dehydration conversion (α), a critical parameter for evaluating photothermal efficiency under irradiation, was quantified viaeqn (1).19 Sintered CaCoSr-15 specimens were subjected to programmed irradiation intervals for thermogravimetric (TG) analysis to determine mass loss percentages (mt). Fig. 2a depicts the correlation between α and irradiation duration (t/min). Under 18 A Xe-lamp irradiation, the α values reached 86.7 wt% and 97.7 wt% after 30 min and 60 min irradiation, respectively. In contrast, the bare Ca(OH)2 control exhibited comparatively negligible α values of 11.4 wt% and 20.4 wt% under identical conditions. This demonstrates a 4.8-fold enhancement in photothermal conversion efficiency for CaCoSr-15 compared to pristine Ca(OH)2. The nanoscale Sr0.4Co2.6O4 component synergistically enhances both the photothermal performance and thermal storage capacity of Ca(OH)2 through a dual-mode photothermal conversion and thermal storage mechanism.
![]() | (1) |
| Qt = αt × Q0 | (2) |
![]() | (3) |
The photothermal storage capacity (Qt) of CaCoSr-15 and Ca(OH)2 under irradiation was quantified through eqn (2) and is visually presented in Fig. 2a. The composite material demonstrated a thermal storage capacity of 905 J g−1 after 60 min irradiation, representing a 3.4-fold enhancement compared to bare Ca(OH)2. The photothermal storage rate (vQ) was determined using eqn (3), with results tabulated in Table 2.21,33 Pure Ca(OH)2 exhibited mean vQ values of 4.93 and 4.42 J g−1 min−1 during 30- and 60-min irradiation periods, respectively. In contrast, CaCoSr-15 achieved substantially higher rates of 28.17 J g−1 min−1 (5.7-fold increase) and 15.08 J g−1 min−1 (3.4-fold increase) under comparable conditions (Table 2). These findings collectively demonstrate that nanoscale Sr0.4Co2.6O4 substantially enhances the photothermal storage kinetics of Ca(OH)2, likely attributable to its elevated photothermal temperature profile and improved energy conversion efficiency.
| v Q (J g−1 min−1) | |||
|---|---|---|---|
| (10 min) | (30 min) | (60 min) | |
| Ca(OH)2 | 8.32 | 4.93 | 4.42 |
| CaCoSr-15 | 64.33 | 28.17 | 15.08 |
The linear regression analysis of TG data for the dehydration process of CaCoSr-15 is presented in Fig. S2. The Arrhenius parameters, activation energy (E) and pre-exponential factor (A), were systematically derived from the slope and intercept of the linear regression lines, respectively. The mean kinetic parameters for CaCoSr-15 dehydration exhibited an activation energy of 166.91 ± 2.06 kJ mol−1 and a pre-exponential factor of 23.20 ± 0.39 (dimensionless). In contrast, bare Ca(OH)2 demonstrated significantly higher kinetic barriers with E = 198.17 ± 9.7 kJ mol−1 and ln
A = 29.40 ± 1.90 (dimensionless).15 The incorporation of 15 wt% Sr0.4Co2.6O4 demonstrated remarkable enhancement in thermal storage performance, reducing the activation energy barrier by 31.3 kJ mol−1 (−15.8%) and decreasing ln
A by 21.1%. This kinetic optimization mechanism confirms the efficacy of Sr0.4Co2.6O4 doping in improving the dehydration kinetics of Ca(OH)2 through reduced activation parameters.19 The temperature-dependent reaction mechanisms for both Ca(OH)2/Sr0.4Co2.6O4 composite and pristine Ca(OH)2 were quantitatively described using the Arrhenius-derived eqn (4) and (5), respectively.
![]() | (4) |
![]() | (5) |
To validate the reliability of the proposed kinetic models, the dehydration conversion (α) was determined from TGA experimental data acquired at a heating rate (β) of 2 °C min−1 and subsequently calculated using eqn (4) and (5). Comparative analysis presented in Fig. S3 demonstrates maximum deviations of 8.6% and 7.4% between experimental and predicted values for Ca(OH)2 and CaCoSr-15 systems, respectively. These residual discrepancies fall within acceptable tolerances for thermochemical analysis, thereby confirming the validity of the kinetic parameters for reactor engineering applications and numerical simulation protocols.
Complementary XPS analysis of the Ca 2p region (Fig. 3b) reveals an upward binding energy shift (ΔBE ≈ 0.6 eV) for the Ca(OH)2–Sr0.4Co2.6O4 composite compared to pristine Ca(OH)2. This electronic perturbation can be attributed to the electron-withdrawing effect induced by the Sr0.4Co2.6O4 component. The observed electronic restructuring phenomena provide mechanistic insight into the enhanced catalytic activity of Sr0.4Co2.6O4 and the reduced activation energy for dehydration reaction pathways (eqn (4) and (5)), as previously reported.13,37
In the high-resolution Co 2p XPS spectrum (Fig. 3c), two primary peaks are observed at binding energies (BEs) of 780.07 and 795.10 eV, corresponding to the Co3+ 2p3/2 and 2p1/2 spin–orbit components in the Co–O environment, respectively. The 15.03 eV splitting between the Co 2p3/2 and Co 2p1/2 spin–orbit components is characteristic of Co3+ oxidation states.38 The weaker peaks at 782.15 and 797.02 eV are assigned to the Co2+ 2p3/2 and 2p1/2 spin–orbit states. Additionally, satellite peaks at ∼789.5 and 804.84 eV are attributed to paramagnetic Co2+ species.39 The peak area ratio (0.31) of Co2+ 2p3/2 to Co3+ 2p3/2 confirms their coexistence within the material, which can be formulated as Sr0.4Co2.6O4.39 Notably, Co2+ ions (d7 electronic configuration) exhibit strong Jahn–Teller activity, promoting lattice defect formation and oxygen vacancy generation.33 Compared to the Co2+ 2p BEs in Co3O4, the corresponding peaks in Sr0.4Co2.6O4 exhibit a significant upshift in binding energy, indicative of electron-withdrawing effects from both Sr2+ doping and OVs in the nanoparticles. In contrast, the Co 2p BEs observed in the CaSrCo-15 composite exhibit a significant upward shift compared to Sr0.4Co2.6O4 (Fig. 3c), indicative of substantial interfacial coupling between Ca(OH)2 and the Co–O framework. This enhanced interaction likely facilitates dehydration through a reduction in activation energy, as delineated in the mechanistic scheme represented by eqn (4) and (5).
The deconvoluted O 1s spectra reveal three distinct binding energy peaks at 529.48, 531.59, and 533.03 eV (Fig. 3d). The primary peak at 529.48 eV corresponds to lattice oxygen species originating from Sr–O and Co–O bonding configurations. The higher-binding-energy peak at 533.03 eV is assignable to adsorbed water molecules on the material surface. Notably, the intermediate peak at 531.59 eV is attributed to OVs, which constitute 52% of the total O 1s peak area. This exceptionally high OV concentration is attributed to the substitution of Sr2+ ions (ionic radius = 1.13 Å) for Co2+ ions (ionic radius = 0.74 Å) within the crystal lattice. The substantial size disparity between these cations induces significant lattice distortion.13,30
Comparative analysis demonstrates that Sr substitution induces a onefold increase in OV content relative to undoped Co3O4 and Cu-doped Co3O4 derivatives. Quantitative evaluation reveals OV populations of 25% and 28% in nano-Co3O4 and Cu0.4Co2.6O4, respectively (Fig. 3e, f, S5, S6, Tables S3 and S4). The enhanced OV concentration (increased 3%) in Cu0.4Co2.6O4 may be systematically elucidated through two synergistic mechanisms: (1) the Jahn–Teller distortion of Cu2+ ions (d9 electronic configuration) that induces structural flexibility, and (2) the ionic radii proximity between Cu2+ (0.72 Å) and Co2+ cations, which effectively mitigates lattice strain during defect formation. In contrast, the considerably larger Sr2+ ions (1.13 Å) generate significant local strain fields, which promote OV formation via amplified metal–oxygen bond polarization.
Notably, the BE values of lattice oxygen in Sr0.4Co2.6O4, Cu0.4Co2.6O4, and Co3O4 nanoparticles were determined to be 529.48 eV, 529.74 eV, and 529.96 eV, respectively (Fig. 3d–f). This systematic increase in BE values follows an opposite trend to the OV content, suggesting that OVs exhibit electron-donor characteristics within these nanoparticle systems (OOV⋯Cox+–Olattice). Consequently, OVs may play a crucial role in the strong absorption of long-wavelength light (Fig. 1a and Table 1). Characterized by the presence of ·O2−, ·O22−, and ·OH− species, OVs may also contribute to the efficient dehydration of hydroxides.29,31
The thermal dehydration of hydroxides proceeds through a radical-chain reaction mechanism. The process is governed by eqn (6), which constitutes the rate-determining step in the dehydration of Ca(OH)2 and exhibits a comparatively elevated activation energy barrier.40 However, this reaction mechanism shifts in the presence of hydroxyl (·OH) and superoxide (·O2−) radicals located at the OVs, where the dehydration reactions (eqn (7)–(9)) are significantly accelerated through radical-mediated pathways with reduced activation energy barriers.
![]() | (6) |
| Ca(OH)2 + ·OH → HO–Ca–O· + H2O | (7) |
| HO–Ca–O· → Ca–O + ·OH | (8) |
| Ca(OH)2 + ·O2− → HO–Ca–O− + ·OOH → CaO + H2O + ·O2− | (9) |
| h+ + OH− → ·OH | (10) |
| h+ + H2O → ·OH + H+ | (11) |
| h+ + ·O2− → O2 | (12) |
| h+ + CoO → Co2O3 | (13) |
| e− + O2 → ·O2− | (14) |
| 2Co2O3 + ·O2− → 2O2 + 4CoO | (15) |
| e− + Co3+–O → Co2+–O | (16) |
The Cu0.4Co2.6O4 system demonstrates hydroxide oxidation-driven oxygen evolution.33 Collectively, illumination induces coupled redox cycles involving Co3+–O ↔ Co2+–O interconversion and O2/·O2− interplay through synergistic interactions between photogenerated carriers and OVs in SrxCo3−xO4 nanoparticles. This creates multiple redox processes with enhanced electron delocalization characteristics (Fig. 4a), which are characteristic of systems exhibiting metallic conductivity features.15 Experimental validation reveals the charge carrier density (n) of Sr0.4Co2.6O4 reaches 1.37 × 1021 cm−3, comparable to quasi-metallic phases such as WO2.83 (6.3 × 1021 cm−3) and Cu1.94S (1 × 1021 cm−3).11,33,41,42 Thereby, a high OV density effectively reduces the bandgap,11 enhances the concentration of photogenerated charge carriers, facilitates redox reactions, and induces pronounced localized surface plasmon resonance (LSPR) effects.
Furthermore, as demonstrated in Table 3, Sr0.4Co2.6O4 nanoparticles exhibit the highest charge carrier density among CeO2–Co3O4,15 Cr2O3–SiO2,14 and Co3O4–Co2SiO4 (ref. 43) systems (SI). Experimental determination also reveals an exceptionally high charge-carrier mobility (µ) of 0.741 cm2 V−1 s−1 for Sr0.4Co2.6O4 nanoparticles, surpassing most reported photothermal materials (Table 3). These quantitative characteristics collectively corroborate the material's superior electrical conductivity (Table 3 and eqn (17)).46
| σ = q·n·µ | (17) |
| Materials | T (K) | Charge carrier density (cm−3) | Conductivity (Ω−1 cm−1) | Mobility (cm2 V−1 S−1) | Hall (cm3 C−1) |
|---|---|---|---|---|---|
| Sr0.4Co2.6O4 | 340 | 1.37 × 1021 | 163 | 0.741 | −4.45 × 10−3 |
| CeO2–Co3O4 (ref. 15) | 340 | 1.12 × 1021 | 113 | 0.630 | −5.58 × 10−3 |
| CrOx–SiO2 (ref. 14) | 352 | 1.04 × 1021 | 110 | 0.594 | −6.25 × 10−3 |
| Co3O4–Co2SiO4 (ref. 43) | 352 | 1.21 × 1021 | 106 | 0.577 | −6.25 × 10−3 |
| Co3O4 (ref. 44) | 773 | 8.26 × 1010 | 10−4–10−2 (ref. 45) | 128 | — |
The Hall coefficient (RH) of Sr0.4Co2.6O4 was measured at −4.45 × 10−3 cm3 C−1, significantly lower than that of n-type silicon (−0.1 cm3 C−1). This negative value signifies that Sr0.4Co2.6O4 is an n-type semiconductor, implying that the doping of Sr2+ ions has led to a transformation from p-type Co3O4. This transformation could be attributed to the high density of OV in the nanoparticles. According to eqn (18), the smallest RH value among the corresponding materials listed in Table 3 suggests that Sr0.4Co2.6O4 nanoparticles have the largest carrier concentration (n), which is in an agreement with the experimentally measured values.
| RH ∝ 1/(n × q) | (18) |
Comparative analysis with state-of-the-art absorber materials confirms Sr0.4Co2.6O4's position as the most conductive substance (163 Ω−1 cm−1) among those documented in Table 3. This enhanced transport behaviour could be attributed to the synergetic effects of OV concentration and redox-active sites within Sr0.4Co2.6O4, which also synergistically enhance broadband absorption across the visible and NIR spectrum (Fig. 1a).15,45 The OV-mediated redox synergy mechanism may additionally account for the diminished light absorption observed in Co3O4 and Cu0.4Co2.6O4 nanoparticles (Fig. 1a and Table 1),41,43,47 which arises from the insufficient OV concentration and limited redox-active components within their nanostructures. Furthermore, comprehensive material characterization confirms that Sr0.4Co2.6O4 exhibits superior photovoltaic performance compared to other listed candidates in Table 4, establishing its exceptional status as an efficient absorber.
| Material | Charge carrier density (cm−3) | Conductivity (Ω−1 cm−1) | Mean Aa | Reference |
|---|---|---|---|---|
| a light absorption A with mean values in visible and NIR region. b FSS, full solar spectrum. | ||||
| (Mn2O3)3CuSiO3 | 3.45 × 1021 | 6.14 × 103 | 0.83 (FSSb) | 33 |
| WO2.83 | 6.30 × 1021 | 2 × 103 | −(FSS) | 12 and 48 |
| TiO0.18N0.86 | 2.40 × 1022 | — | −0.65 (FSS) | 49 |
| Fe0.92S2 | 1.91 × 1022 | — | −(FSS) | 50 |
| Cu1.8S | ∼1021 | — | 0.8 (FSS) | 12 and 51 |
| LaCoOx | 1.30 × 1021 | 163 | 1.12 (FSS) | 52 |
| CrOx–SiO2 | 1.04 × 1021 | 110 | 1.46 (FSS) | 14 |
| Co2SiO4–SiO2 | 1.20 × 1021 | 108 | 1.27 (FSS) | 28 |
| Co3O4–Co2SiO4 | 1.36 × 1021 | 106 | 1.38 (FSS) | 43 |
| Co3O4–Co3(PO4)2 | 1.20 × 1021 | 101 | 1.17 (FSS) | 13 |
| CeO2–Co3O4 | 1.12 × 1021 | 113 | 0.86 (FSS) | 15 |
| Sr0.4Co2.6O4 | 1.37 × 1021 | 163 | 1.36 (FSS) | This paper |
The A1g vibrational mode of Cu0.4Co2.6O4 exhibits a pronounced redshift, with spectral positions shifting from 651 to 642 cm−1. This spectral shift suggests a weakening of Co–O bond strength induced by Cu2+ doping. The observed bond strength reduction is primarily ascribed to significant lattice geometric reorganization in Cu0.4Co2.6O4, which may originate from the pronounced Jahn–Teller distortion of Cu2+ (d9) ions.33,54 This modification of metal–oxygen bonding interactions facilitates charge carrier mobility enhancement and redox reaction kinetics, thereby promoting increased charge carrier concentration and improved near-infrared light absorption efficiency, as experimentally demonstrated in Fig. 1a.53
Additionally, energy-dispersive X-ray spectroscopy (EDS) and scanning electron microscopy (SEM) elemental mapping analyses demonstrate a homogeneous distribution of Co and Sr elements within Sr0.4Co2.6O4 nanoparticles, exhibiting atomic fractions of 87.14% and 12.86% respectively (Fig. 6b–d). These experimental values show statistical equivalence to the theoretical atomic ratio (86.7
:
13.3) of elemental composition in this compound.
Systematic characterization via SEM reveals the micromorphology of as-synthesized Sr0.4Co2.6O4, featuring distinct polyhedral nanocrystalline structures coexisting with submicron-scale particles exhibiting reduced crystallinity. The material displays a well-developed porous architecture with varied pore dimensions (Fig. 7a and b).59 This distinctive nanoarchitecture significantly enhances solar absorption capabilities while improving photothermal conversion efficiency through localized surface plasmon resonance (LSPR)-mediated thermalization mechanisms.6,60
The micromorphology and compositional characteristics of Sr0.4Co2.6O4 were systematically characterized through transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) analyses. The interplanar spacings of 0.206 nm, 0.249 nm, and 0.638 nm corresponded to the (223), (300), and (101) crystallographic planes of Sr0.4Co2.6O4, respectively (Fig. 7c, d and S8). These lattice fringes demonstrated a good correspondence with the diffraction peaks at 2θ = 32.65°, 43.96°, and 12.89° in the X-ray diffraction spectrum (Fig. 6a, 7c and d), confirming phase consistency.
The synthesized Ca(OH)2–Sr0.4Co2.6O4 composite exhibited distinctive hierarchical features as revealed by scanning electron microscopy (SEM) imaging (Fig. 7e and f). The Sr0.4Co2.6O4 phase effectively encapsulated Ca(OH)2 particles with a porous nanostructure, which could mitigate particle agglomeration during thermal energy storage/release processes.19,21 Notably, the shell architecture contained interconnected pore networks derived from nanoparticle and nanorod assemblies, which would facilitate enhanced mass transport during charge/discharge cycles.61 The EDS and SEM elemental mapping analyses for the Ca(OH)2–Sr0.4Co2.6O4 composite are presented in Fig. S9. These analyses clearly demonstrate a homogeneous distribution of cobalt (Co) and strontium (Sr) elements within the composite. The corresponding mass fractions are determined to be 81.65% for Co and 18.35% for Sr, as detailed in Table S6. Importantly, these experimentally obtained values exhibit excellent agreement with the theoretical mass ratio (81.40
:
18.60) of the elemental composition in this composite material.
The TEM and HRTEM images of Ca(OH)2–15 wt% Sr0.4Co2.6O4 composite further corroborate its porous architecture, the surface phase composition of Sr0.4Co2.6O4, and the presence of numerous OVs, which are clearly delineated by circular markers (Fig. 8).
![]() | ||
| Fig. 8 (a) TEM image of Ca(OH)2–15 wt% Sr0.4Co2.6O4 composite. (b) HRTEM image of Ca(OH)2–15 wt% Sr0.4Co2.6O4 composite with circular markers for OVs. | ||
Valence band (VB) edge positions (EVB) were determined through high-resolution XPS valence band spectra, revealing energy levels at 0.075, 0.43, −0.11, −0.01, and 0.54 eV for Sr0.4Co2.6O4, Cu0.4Co2.6O4, Sr0.7Co2.3O4, Sr0.3Co2.7O4, and Co3O4, respectively (Fig. 9b–f). The strontium-doped derivatives exhibit considerably higher EVB values than the undoped Co3O4, with these negative energy levels corroborating the metallic character of the materials. This elevated EVBs are possibly resulted from the OVs, leading to narrowed band gaps. This electronic feature aligns with the enhanced free carrier concentration (FCC) observed in Sr0.4Co2.6O4 through Hall effect measurements.
Conduction band (CB) positions (ECB) were calculated using eqn (19), yielding respective values of −0.275, 0.03, −0.51, −0.44, and 0.03 eV for Sr0.4Co2.6O4, Cu0.4Co2.6O4, Sr0.7Co2.3O4, Sr0.3Co2.7O4, and Co3O4 (Fig. 10b). The comparative analysis indicates that OV formation simultaneously induces band gap narrowing and valence band elevation through two synergistic mechanisms: (1) increased free carrier density elevating the Fermi level, and (2) defect-induced electronic structure modification. These findings establish a clear correlation between OV concentration and the photothermal and optoelectronic properties critical for photothermal and photocatalytic applications.
| ECB = EVB − Eg | (19) |
![]() | ||
| Fig. 10 (a) Ultraviolet photoelectron spectroscopy (UPS) spectra of Sr0.4Co2.6O4. (b) Band structure of Sr0.4Co2.6O4, Sr0.3Co2.7O4, Sr0.7Co2.3O4, Cu0.4Co2.6O4, and Co3O4. | ||
According to band theory, electron transfer dynamics exhibit a direct dependence on the electronic work function (Φ). The work functions of these materials were investigated through ultraviolet photoelectron spectroscopy (UPS) measurements.14,62 Experimental determination of the secondary electron cut-off energy for Sr0.4Co2.6O4 yielded 17.21 eV (Fig. 10a). Through energy band analysis, the work function was calculated as −4.01 eV by subtracting the He I excitation energy (21.22 eV) from the secondary electron cut-off energy. This establishes the Fermi level (EF) at 4.01 eV (where EF = −Φ), significantly below the valence band energy (EVB = 0.075 eV) of Sr0.4Co2.6O4 (Fig. 10b). The observed Fermi level positioning below the valence band indicates that Sr2+ doping induces a downward shift of the Fermi level and metallic property.63
This phenomenon aligns with established semiconductor physics principles, where p-type dopants exhibiting reduced Fermi levels (EF < EVB) demonstrate metallic conduction characteristics. Such behaviour arises from enhanced hole carrier concentration due to dopant-induced band structure modification, manifesting as increased free charge carrier density and improved electrical conductivity.63 The experimental observations corroborate this mechanism, showing coincident high free carrier concentration (1.37 × 1021 cm−3), substantial oxygen vacancy formation (52%), and elevated electrical conductivity (163 Ω−1 cm−1, Table 3) in Sr0.4Co2.6O4 nanoparticles.
![]() | ||
| Fig. 11 (a) Photocurrent response of Co3O4, Cu0.4Co2.6O4, and Sr0.4Co2.6O4. (b) Photoluminescence spectra. (c) Photoluminescence decaying characteristics. | ||
Photoluminescence (PL) spectroscopy demonstrates that the PL intensity follows a descending order of Co3O4 > Cu0.4Co2.6O4 > Sr0.4Co2.6O4, with all exhibiting less than 1% intensity relative to SiO2 (Fig. 11b).14 This inverse correlation with OV content suggests that the suppressed radiative recombination rate is inversely proportional to OV concentration. The sub-1% PL intensity confirms that non-radiative recombination dominates the energy relaxation pathway of photoexcited carriers, thereby enhancing photothermal conversion efficiency.14 The highest PL intensity of the Co3O4 occurs in the wavelength range of 414–482 nm. The average PL intensity within this region was calculated to be 1.26 a. u. The photothermal conversion efficiency of this nano Co3O4 was determined to be 47.7%. For comparison, the photothermal conversion efficiency and PL intensity of Co3O4, Cu0.4Co2.6O4, and Sr0.4Co2.6O4 were determined and are summarized in Table S7. Notably, Sr0.4Co2.6O4 exhibits the highest photothermal performance (86.2%) among these oxides.
Time-resolved photoluminescence analysis (Fig. 11c) reveals distinct carrier relaxation dynamics through two exponential components (τ1 and τ2), corresponding to different recombination mechanisms. The prolonged τ2 component indicates hindered energy dissipation from excited electrons due to combined contributions from lattice defects and redox reaction.64 This two-component decay profile provides critical insights into the competing radiative and non-radiative pathways governing carrier dynamics in these materials.
The fluorescence decay dynamics of Co3O4, Cu0.4Co2.6O4, and Sr0.4Co2.6O4 were systematically investigated through time-resolved photoluminescence (TRPL) spectroscopy, with the corresponding τ values tabulated in Table 5. The considerably shortened fluorescence τ across all samples suggest that non-radiative recombination processes predominantly govern the energy dissipation pathways of photoexcited electrons. Notably, Sr0.4Co2.6O4 exhibits the shortest τ2 component and average τ value alongside the weakest PL intensity, which can be attributed to synergistic quenching mechanisms involving OVs populated by paramagnetic species (O2 and O2−), redox interactions between photoexcited carriers and the Co3O4 matrix, and its elevated photothermal temperature.65 Intriguingly, the observed decreasing sequence of τ2 (7.56 → 6.84 → 6.28 ns) and τavg values across Co3O4, Cu0.4Co2.6O4, and Sr0.4Co2.6O4 contrasts with the OV concentration gradient revealed by X-ray photoelectron spectroscopy (XPS) analysis (Tables S3 and S4). This paradoxical observation implies that OV introduction not only accelerates radiative decay through enhanced electron-phonon coupling but also establishes a linear correlation between OV concentration and fluorescence decay rate constant. Such behaviour contradicts conventional understanding where OV defects typically prolong radiative lifetimes.65,66 Our findings demonstrate that OV engineering simultaneously optimizes photothermal conversion efficiency and photocurrent density by strategically modulating both PL intensity and carrier relaxation dynamics.
| Material | τ 1/ns (rel%) | τ 2/ns (rel%) | τ avg/ns |
|---|---|---|---|
| Co3O4 | 1.35 (48.9) | 7.56 (51.1) | 4.52 |
| Cu0.4Co2.6O4 | 1.39 (43.6) | 6.84 (56.4) | 4.44 |
| Sr0.4Co2.6O4 | 1.31 (42.9) | 6.28 (57.1) | 4.14 |
| This journal is © The Royal Society of Chemistry 2026 |