Santosh V.
Mohite‡
a,
Artavazd
Kirakosyan‡
b,
Kwangchan
An
a,
Yeongseok
Shim
a,
Jihoon
Choi
*b and
Yeonho
Kim
*a
aDepartment of Applied Chemistry, Konkuk University, Chungju 27478, Republic of Korea. E-mail: yeonho@kku.ac.kr
bDepartment of Materials Science and Engineering, Chungnam National University, Daejeon 34134, Republic of Korea. E-mail: jihoonc@cnu.ac.kr
First published on 23rd June 2025
Size-controlled cuprous oxide-based nanoparticles (NPs) are promising materials for enhancing visible-light-driven photocatalytic hydrogen production by increasing the number of Cu+ surface active sites. This study investigates the role of molecular additives in the growth of Cu/Cu2O NPs on mesoporous silica (m-SiO2) templates. The molecular additives cetyltrimethylammonium bromide (CTAB), ascorbic acid (AA), and citric acid (CA) are analyzed for their ability to modify the zeta potential of m-SiO2, facilitating the adsorption of Cu2+ ions. The modified surface effectively controls the interaction between Cu2+ ions and the m-SiO2 surface through the influence of molecular additives. The CTAB system facilitates a rapid nanoparticle (NP) growth and significant aggregation, thereby promoting the adsorption of Cu species and the subsequent formation of larger NPs. In contrast, CA provides better control over the formation of NPs, preventing aggregation through Cu2+ chelation and stabilizing the particles within the mesoporous voids of silica. Furthermore, the intensity ratio of metallic Cu to Cu2O has the lowest value of 0.47 in the CA system, indicating a higher Cu2O content compared to the CTAB and AA systems. CTAB and AA favor metallic Cu formation, while CA stabilizes Cu+ and promotes Cu2O phase growth. As a result, the CA system achieves a 5-fold increase in the hydrogen production rate under visible light compared to the CTAB system. These findings highlight the critical role of molecular additives in tailoring NP growth and enhancing photocatalytic performance.
Cuδ+ active sites are formed in copper-based catalysts through phase composition, which distinguishes their chemical and electrical properties, enabling the exploration of the active surface for effective reduction reactions.10–13 The Cu+ site is the active center during photocatalytic H2 production that interacts with H+ ions to produce molecular hydrogen. However, excessive oxidation of Cu+ to Cu2+ on the surface of catalysts can reduce the availability of active sites.14,15 This is accomplished by structural control and multifunctional involvement in catalytic reactions, enhancing the catalyst's performance at the molecular level. For example, Qi Wu et al. reported that Cu(I) sites are generated in situ by reducing Cu(II) salts using AA.16 These generated Cu(I) sites are adsorbed onto the polymer surface, where they interact with nitrogen-rich triazole groups. These interactions stabilize Cu(I) and facilitate the formation of Cu2O nanocrystals under weakly acidic conditions. However, these active sites become unstable during photocatalytic reactions, resulting in reduced efficiency and performance over time. Instability can be induced by surface degradation or photocorrosion, which diminishes the catalyst's long-term efficiency.
Photocorrosion of Cu2O occurs due to self-photooxidation during photocatalytic reactions.17 This critical factor greatly influences and diminishes active site density, which can significantly reduce photocatalytic hydrogen performance. Various studies, for example, constructing core–shell structures enwrapped with graphene oxide, carbon-encapsulated nanostructures, or structures supported with metal–organic frameworks (MOFs), have been conducted to minimize the photocorrosion of Cu2O and preserve the surface active sites.18–20 Wang et al. reported that Zn-doped Cu2O supported by graphdiyne (GDY) provides abundant active sites and redistribution of surface charges.21 The formation of an S-scheme heterojunction between Zn–Cu2O and GDY created an internal electric field that suppressed electron–hole recombination. However, stability tests conducted with eosin Y (EY) as a sensitizer revealed that the H2 production efficiency decreased over multiple reaction cycles due to photocorrosion and the degradation of EY molecules under prolonged light exposure. The researchers examined the porous nature of MOFs, which provide a protective environment for the effective stabilization of Cu2O NPs, with their high surface area and tunable structures. The coordination between the metal sites in the MOFs and the Cu2O NPs creates strong interactions that are advantageous for heterojunction formation, which prevents charge carrier recombination by creating internal electric fields and minimizes photocorrosion.22 Nevertheless, binder-free supported Cu2O NPs are likely to aggregate, which results in limited numbers of active sites and surface areas. In this context, stabilized Cu–Cu2O NPs supported with SiO2 reduce aggregation and enhance the dispersion of Cu2O NPs.23,24 The mesoporous structure of SiO2 not only stabilizes Cu2O-based NPs but also facilitates efficient reactant diffusion. Moreover, m-SiO2 has a negatively charged surface that strongly interacts with Cu+ ions for adsorption, facilitating nucleation sites for the growth of NPs.25,26 This interaction can be controlled using molecular additives such as CTAB or AA.27,28 Molecular additives used during the synthesis process can significantly affect the formation and stabilization of active sites on the surface of Cu-related phases, which are supportive of SiO2.29,30 However, these molecular additives vary the reactive environment of the negatively charged mesoporous SiO2 surface to control the surface interaction of Cu ions. Also, it can be beneficial for the reduction of Cu ions in copper phases. These molecular additives have been shown to modify the phase distribution by influencing the nucleation and growth of copper phases.
This study investigated the influence of different molecular additives such as CTAB, AA, and CA on the preparation of the copper phase composition on m-SiO2. The catalytic activity of these NPs was evaluated for photocatalytic H2 production, focusing on understanding how the additives influenced the growth and aggregation of Cu species on the m-SiO2 surface. HR-TEM analysis was used to examine the surface morphology and the role of these molecular additives in controlling the distribution of Cu species within the mesoporous voids of SiO2. Furthermore, XRD analysis was employed to identify the metallic Cu and Cu2O nanocrystalline phases formed in the nanocomposite. In addition, the XRD intensity ratios were analyzed to confirm the role of molecular additives in the synthesis of Cu-related crystalline phases. The results demonstrated that molecular additives significantly affect the growth rate of Cu species, aggregation behavior, and the formation of crystalline phases on the positively charged m-SiO2 surface. Among the additives, CA was found to be particularly effective in optimizing and controlling electrostatic interactions with the mesoporous SiO2 surface, ensuring uniform distribution and stable formation of Cu/Cu2O NPs.
![]() | ||
Fig. 1 Schematic representation of the preparation of Cu/Cu2O NPs on mesoporous SiO2 using various molecular additives. |
Sample | Solvent | Molecular additive | Zeta potential (mV) |
---|---|---|---|
m-SiO2 | DMSO | None | −44.70 ± 0.15 |
CTAB | −38.20 ± 1.40 | ||
AA | −33.97 ± 0.21 | ||
CA | −28.40 ± 0.22 |
HR-TEM images were used to examine the loading behavior of the Cu/Cu2O NPs on the surface of m-SiO2 using different molecular additives (Fig. 2a–c). The TEM images reveal significant differences in aggregation behaviour among the samples. Notably, the aggregation of Cu species on the m-SiO2 NP surface is most pronounced in the CTAB-assisted system compared to the CA-assisted and additive-free system. This increased aggregation of NPs in the system is attributed to the strong surface adsorption of Cu ions, which facilitates the formation of nucleation sites for Cu2+ species in solution. Because the zeta potential values are below the isoelectric point, the CTAB on the negatively charged m-SiO2 results in a less negative zeta potential value compared to that on bare m-SiO2, but the surface is not fully neutralized.32 This modified surface facilitates control of the adsorption and aggregation of Cu species on the SiO2 surface. Furthermore, during hydrogenation treatments, decomposition of CTAB occurs, leading to the formation of aggregated Cu/Cu2O NPs. The CA-assisted system shows reduced NP aggregation due to more controlled Cu2+ adsorption. As a weak molecular acid, CA interacts with m-SiO2 through its three carboxylate groups, lowering the zeta potential from −38.20 to −28.40 mV. Compared to AA and CTAB, CA achieves higher surface coverage and better charge neutralization. Furthermore, CA chelates Cu2+ ions, enabling uniform dispersion and limiting nucleation sites. N. Balighieh et al. reported that chelation involves coordinating metal ions with silicon to form coordination complexes.33 Consequently, a metallic citric complex formed at the atomic scale.34 According to P. Sukpanish et al., the presence of carboxyl and hydroxyl groups on the surface of this compound facilitates the formation of metallic–citrate complexes through coordination with metal ions.35 This suggests that the CA molecule can act as a bridge surrounding the SiO2 surface via Cu2+ ions and control the extensive aggregation of NPs. This behavior promotes more sites for homogeneous nucleation and controlled growth of Cu species, preventing excessive aggregation on the m-SiO2 surface. In the absence of molecular additives, the negatively charged m-SiO2 surface induces electrostatic attraction with Cu2+ ions, leading to rapid nucleation and subsequent growth of Cu species. However, this process lacks the controlled distribution observed in the CA-assisted system. HR-TEM images of the AA-assisted system are shown in Fig. S3.† Therefore, it is inferred that CA has the optimum reaction rate for growing Cu species on m-SiO2 and prevents rapid aggregation of Cu species on the surface of m-SiO2. HAADF-STEM analysis also supports these findings, demonstrating a similar trend in the aggregation and distribution of Cu species (Fig. 2d–f). Elemental mapping images further confirm the uniformity and distribution of Cu elements in the CA-assisted system compared to the others (Fig. 2j–l).
Fig. 3 shows the XRD patterns of the photocatalysts prepared with different molecular additives. All samples exhibit characteristic peaks corresponding to cubic metallic Cu and Cu2O phases, which clearly match JCPDS cards 003-1005 and 005-0667, respectively. Variations in the diffraction intensities of these phases were observed. These exposed phases depend on the molecular additive used in the synthesis of NPs. The intensity ratio of the major diffraction peak of metallic Cu to that of Cu2O was calculated to determine the dominant Cu phase in each sample (Fig. 3b). The intensity ratios for Cu/Cu2O–CTAB, Cu/Cu2O–AA, Cu/Cu2O, and Cu/Cu2O–CA were 4.52, 4.22, 0.48, and 0.47, respectively. These results suggest that molecular additives such as CTAB and AA favor the formation of metallic Cu, whereas the CA system significantly suppresses metallic Cu and enhances the Cu2O phase. This suggests that CA supports the stabilization of oxidized Cu2O phases through its chelation capability, which suppresses complete reduction to metallic Cu. In contrast, AA primarily acts as a reducing agent, promoting rapid conversion of Cu2+ to Cu+ and Cu0. Thus, CA treatment yields a higher population of Cu+ active sites than AA, as shown in Fig. 3a. This is beneficial for photocatalytic H2 generation. Further insights were provided by FT-IR analysis, as shown in Fig. 3c, which confirmed the presence of functional groups associated with the Cu2O structure. The spectra exhibited Cu–O stretching vibrations at around 616 and 666 cm−1, consistent with the Cu2O lattice structure.36 Interestingly, minor peaks corresponding to Cu2S were also observed in the XRD patterns. These peaks are attributed to the synthesis conditions, particularly the use of DMSO as a solvent. At 100 °C, the highly viscous nature of the reaction mixture due to DMSO prevented complete solvent evaporation. This concentrated environment, enriched with molecular additives and Cu2+ complexes, facilitated the formation of Cu–DMSO complexes.37 The thermal decomposition of these complexes resulted in crystalline Cu2S.
![]() | ||
Fig. 3 (a) XRD patterns and (b) intensity ratios of the major diffraction planes of metallic Cu to Cu2O. (c) FT-IR spectra of Cu/Cu2O, Cu/Cu2O–CTAB, Cu/Cu2O–AA, and Cu/Cu2O–CA. |
The surface oxidation states and bonding characteristics of the Cu species were analyzed through deconvoluted XPS spectra of Cu 2p, as shown in Fig. 4a. Peaks at 932.22 and 934.00 eV are attributed to the Cu0/Cu+ and Cu2+ oxidation states, respectively.38 The presence of strong satellite peaks further confirms the Cu2+ oxidation state. These results indicate that the synthesized NPs consist of a mixture of Cu2O, metallic Cu, and CuO, suggesting that the molecular additives used during synthesis facilitated the formation of mixed oxidation states (Fig. 4a). These findings suggest that molecular additives promote the formation of mixed oxidation states in cuprous oxide-based NPs.
![]() | ||
Fig. 4 (a) Cu 2p and (b) O 1s XPS spectra of Cu/Cu2O, Cu/Cu2O–CTAB, Cu/Cu2O–AA, and Cu/Cu2O–CA. (c) TRPL spectrum of Cu/Cu2O–CA. (d) Average lifetimes of Cu/Cu2O–CTAB, Cu/Cu2O–AA, and Cu/Cu2O–CA. |
The influence of molecular additives on the surface chemical environment is evident from the shifts in XPS peak positions. The spin–orbit splitting between the Cu 2p3/2 and Cu 2p1/2 peaks was approximately 19.8 eV. However, variations in binding energy shifts were observed, reflecting differences in oxidation states and surface environments. For Cu/Cu2O–AA, lower shifts in binding energy indicate a higher proportion of metallic Cu, consistent with the reducing nature of AA. Conversely, a positive binding energy shift was observed in Cu/Cu2O–CA, which promotes the stabilization of the Cu2O phase. Compared to the AA and CTAB systems, the CA-treated catalysts and those prepared without reducing agents exhibited a higher Cu2O content. In addition, the peak at 935.85 eV in the deconvoluted XPS spectra was attributed to the hydroxyl groups bonded to Cu species. The bonding of oxygen to Cu in the lattice was confirmed through O 1s XPS peaks (Fig. 4b). The peak at 531.1 eV corresponds to oxygen atoms integrated into the crystal lattice of the oxide. In comparison, the strong peak at 532.7 eV is associated with the m-SiO2-support material. Further insights were obtained from the XPS spectra of Si 2p (Fig. S4†) and C 1s (Fig. S5†). The C 1s peaks likely originate from molecular additives or residual carbon left after incomplete decomposition during synthesis. These carbon functional groups were also confirmed in the FT-IR spectra (Fig. 3c), with a band at 1650 cm−1 indicating residual molecular compounds in the catalyst.
Fig. 4c and S6 and S7† show the TRPL decay profiles of the Cu/Cu2O–CA, Cu/Cu2O–CTAB, and Cu/Cu2O–AA catalysts. The decay profiles were analyzed using a tri-exponential decay fitting model,39 providing lifetime components of τ1, τ2, and τ3, which correspond to fast, intermediate, and slow recombination processes, respectively. These components are summarized in Table S1.† The τ1 is attributed to non-radiative recombination involving surface trap states, while τ2 and τ3 reflect charge carriers trapped by shallow trap states caused by structural defects and radiative recombination of charge carriers.40,41 Among the systems, Cu/Cu2O–CA exhibits a higher τ2 value than Cu/Cu2O–AA but lower than Cu/Cu2O–CTAB, suggesting efficient interfacial charge transfer in the CA-assisted composite NPs. This trend is consistent with the mean lifetime (τavg), as shown in Fig. 4d. The τavg values for Cu/Cu2O–CTAB, Cu/Cu2O–AA, and Cu/Cu2O–CA are 4.14, 3.30, and 3.43 ns, respectively. The higher τavg value in the CTAB system may be attributed to the formation of a higher proportion of the Cu2S phase, which can influence recombination processes. However, the slightly increased τavg value in the CA system (1.04 times higher than that of Cu/Cu2O–AA) indicates improved charge transfer efficiency.
Fig. 5a–c show the Mott–Schottky (MS) plots of Cu/Cu2O–CTAB, Cu/Cu2O–AA, and Cu/Cu2O–CA. All samples exhibit a negative slope in the linear region of the MS plots, confirming their p-type conductivity. Notably, this conductivity remains unaffected by the use of different molecular additives. The flat band potential (Vfb) was determined by extrapolating the linear portion of the MS plots at 1/C2 = 0. The Vfb values for Cu/Cu2O–CTAB, Cu/Cu2O–AA, and Cu/Cu2O–CA were found to be 1.48, 1.54, and 1.33 V vs. RHE, respectively. The observed variation in the Vfb value is attributed to the presence of mixed Cu-related crystalline phases. The acceptor concentration (NA) was calculated using the formula NA = 2/(q·ε·ε0 × slope). The NA values were 5.80 × 1020 cm−3, 8.86 × 1020 cm−3, and 1.15 × 1021 cm−3 for Cu/Cu2O–CTAB, Cu/Cu2O–AA, and Cu/Cu2O–CA, respectively (Fig. 5d). Cu/Cu2O–CA demonstrates the highest NA value, indicating improved charge carrier mobility and charge carrier separation. This was confirmed by the photoelectrochemical (PEC) response (Fig. 5e). The superior performance of the CA system is likely due to the abundance of surface-active sites on m-SiO2, which effectively support NP loading. In contrast, the CTAB-assisted system exhibited significant NP aggregation, minimizing the exposed surface area and reducing the accessibility for active sites. Furthermore, residual CTAB within the NPs, retained during synthesis, decomposes between 200 and 300 °C.42 The retained residual of CTAB obstructed the surface-active sites, hindering their accessibility for catalytic reactions. This limitation can adversely affect the efficiency of surface-related reactions by reducing the number of available active sites and diminishing the overall performance. The charge transport behavior of Cu/Cu2O–CA under light and dark conditions was analyzed through electrochemical impedance spectroscopy (EIS), as shown in Fig. 5f. The charge transfer resistance (Rct) was determined by fitting the EIS data to an equivalent circuit model.39 For Cu/Cu2O–CA, the Rct value decreased from 164 × 103 Ω under dark conditions to 114 × 103 Ω under light irradiation. This significant reduction suggests efficient utilization of photogenerated charges for the reduction of H+ ions, driving visible light-induced hydrogen production.
The H2 production activity of the photocatalysts synthesized at different temperatures (200, 300, and 400 °C) was evaluated under visible light irradiation. As shown in Fig. S8,† the sample prepared at 300 °C exhibited the highest H2 evolution rate. This temperature optimally balances the decomposition of organic ligands and the stabilization of catalytically active Cu+ species on the m-SiO2 surface. Fig. 6a and b show the H2 production activities of catalysts prepared using CA, AA, and CTAB. An increase in H2 production was observed after one hour of illumination. This may be attributed to the activation of the photocatalyst surface through the formation of a secondary phase. These results are consistent with previous studies reporting metal sulfide phase formation during H2 production.43 The H2 production rates for Cu/Cu2O–CTAB, Cu/Cu2O–AA, and Cu/Cu2O–CA were determined to be 38.33, 86.49, and 135.81 μmol g−1 h−1, respectively. These results clearly indicate that Cu/Cu2O–CA exhibits superior photocatalytic activity compared to the other systems. The enhanced performance of Cu/Cu2O–CA can be attributed to the stabilization of nanosized Cu species within the voids of the m-SiO2 surface, which prevents NP aggregation and increases the active surface area for the H2 production reaction. Moreover, the higher availability of Cu+ sites in the CA system further promotes efficient charge separation and facilitates proton reduction to hydrogen, thereby enhancing the overall photocatalytic activity.
In contrast, the Cu/Cu2O–CTAB system displayed rapid aggregation of NPs, significantly reducing the number of active sites available for catalysis and, consequently, lowering H2 production. To further evaluate the performance of Cu/Cu2O–CA, its photocatalytic H2 production rate was compared with that of commercial Cu2O NPs. Remarkably, Cu/Cu2O–CA exhibited a 9.1-fold higher H2 production rate than the commercial Cu2O NPs, underscoring the effectiveness of the synthesized catalyst (Fig. 6b). Moreover, the effect of varying CA concentrations on H2 production activity was investigated (Fig. S9†). The results revealed that the catalytic performance improved with increasing CA concentration, reaching an optimal point. This enhancement can be attributed to the dual functionality of CA as both a structure-directing agent and a metal-complexing ligand, which promotes uniform dispersion of copper species and better control over NP morphology. Furthermore, the H2 production activity of optimized Cu/Cu2O–AA was tested in the presence of different sacrificial agents, including 10 v/v% TEOA, methanol, and glycerin, as shown in Fig. 6c. The results demonstrated a linear, time-dependent increase in H2 production activity. However, the rates were lower compared to those obtained using a Na2S/Na2SO3 scavenger. These findings suggest that the reaction of the catalyst surface with S2− ions in the Na2S/Na2SO3 system contributes to the formation of a secondary phase, enhancing H2 production efficiency. The order of H2 production rates (Fig. 6d) for various scavengers with Cu/Cu2O–AA was as follows: Na2S/Na2SO3 > TEOA > methanol > glycerin. Notably, the Na2S/Na2SO3 scavenger exhibited the highest H2 production rate, 12.8-, 15.5-, and 458.8-fold higher than those of TEOA, methanol, and glycerin, respectively. The stability and reusability of the Cu/Cu2O–CA catalyst were evaluated through ten consecutive photocatalytic H2 production cycles (Fig. 6e). Interestingly, the activity in the second cycle increased by 1.26-fold compared to the first run, as shown in Fig. 6e, possibly due to surface reconstruction or activation. A slight decline in activity was observed in later cycles, which correlates with structural and compositional changes in the catalyst. XRD analysis (Fig. S10†) revealed a phase transformation from Cu2O to Cu2S during the reaction. The XPS spectra (Fig. S11†) exhibited a shift of Cu 2p peaks toward lower binding energies, which may be attributed to surface chemical changes induced by sulfur species with copper. This interaction is responsible for the observed increase in sulfur species on the catalyst surface during H2 production, as clearly shown in the EDS mapping images (Fig. S12 and S13†). The elemental maps reveal a denser and more homogeneous distribution of sulfur after the reaction. Quantitative EDS data, as summarized in Table S2,† indicate a significant increase in the atomic percentage of sulfur. Elemental analysis (Table S3†) confirms this trend, revealing a threefold increase in sulfur content after the photocatalytic reaction. These findings strongly suggest the in situ formation of Cu2S species during H2 evolution.
The charge transfer mechanism in the Cu/Cu2O/Cu2S heterojunction for photocatalytic H2 production is schematically illustrated in Fig. 7. The proposed mechanism is based on the energy band positions and their role in the transfer of photogenerated electron–hole pairs. Metallic Cu has a work function of 4.20–4.65 eV.44 Cu2O, a p-type semiconductor, has its Fermi level above the valence band,45 with an electron affinity of approximately 3.2 eV (ref. 46) and a work function of 4.80–5.00 eV.47,48 The difference in the work functions of Cu and Cu2O creates a barrier height at the Cu2O/Cu interface,49 forming an electron-rich region. This induces electron transfer between the materials until equilibrium is reached. This leads to the formation of band bending due to dissimilar work functions, which creates a built-in electric field.50 The built-in electric field within the Cu/Cu2O/Cu2S heterojunction plays a crucial role in enhancing charge separation and suppressing charge recombination. The photogenerated electrons and holes possess thermodynamically sufficient potentials to drive water splitting.51 In particular, Cu+ species act as more active surface sites for redox reactions. The use of CA as a molecular additive stabilizes a higher proportion of Cu+ species, as evidenced by XRD analysis, where the Cu2O phase was the dominant component in the CA-prepared catalyst. In contrast, the CTAB and AA systems showed higher metallic Cu content. These compositional variations between metallic Cu and Cu2O can alter the electronic properties of the material. They affect the alignment of band positions, the Fermi level, and the density of states at the interface.52 Therefore, the CA additive promotes the formation and retention of Cu2O, which has a conduction band edge sufficiently negative to reduce protons. Suppaso et al. have demonstrated that Cu-based heterojunctions benefit from such favorable energy band alignment, which facilitates photogenerated electron transfer to active sites, enabling efficient visible-light-driven water splitting.12 As a result, favorable energy band alignment directly influences the charge carrier separation efficiency and the overall photocatalytic performance of the material. Upon photon absorption by Cu2O, electrons are excited from its valence band (VB) to its conduction band (CB). Simultaneously, the built-in electric field at the heterojunction interface facilitates the migration of holes toward the Cu2S side. The photogenerated electrons in the CB of Cu2O are subsequently transferred to the Cu layer and react with H+ ions, resulting in molecular hydrogen.53 During H2 production, the partial transformation of Cu2O into Cu2S was confirmed by elemental analysis, which further supports the formation of a heterostructure that aids in charge separation. Such multi-phase heterojunctions highlight the synergistic effects of compositional tuning, interface engineering, and additive-assisted phase control in enhancing the photocatalytic H2 production performance of the Cu/Cu2O–CA system, as shown in Fig. 6a.
![]() | ||
Fig. 7 Charge transfer and separation mechanism through a built-in electric field over the surface of Cu/Cu2O–CA supported with Cu2S for H2 production. |
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5nr02191j |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |