Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Insights into bifunctional active sites of Pt–MoO3/TiO2 catalysts enabling selective hydrogenation of an amino acid

Yundao Jing a, Xiaohu Gea, Rui Songa, Ningchao Zhua, Jinquan Minga, Nihong Anb, Yueqiang Cao*a, Gang Qiana, Xuezhi Duan*a and Xinggui Zhoua
aState Key Laboratory of Chemical Engineering and Low-Carbon Technology, School of Chemical Engineering, East China University of Science and Technology, 130 Meilong Road, Shanghai 200237, China. E-mail: yqcao@ecust.edu.cn; xzduan@ecust.edu.cn
bSino-Platinum Industrial Catalyst (Yunnan) Co., Ltd., 988 Keji Road, Kunming 650106, Yunnan, China

Received 31st July 2025 , Accepted 7th October 2025

First published on 9th October 2025


Abstract

Selective hydrogenation of amino acids to amino alcohols is a valuable transformation in the synthesis of pharmaceuticals, fine chemicals, and chiral building blocks. However, achieving high activity and selectivity under mild conditions remains challenging due to the need for simultaneous hydrogen activation and substrate coordination. Here, we report a series of Pt–MoO3 bifunctional catalysts for the hydrogenation of L-alanine (Ala) to alaninol (AlaOH), with a focus on tuning metal–oxide synergy. Structural and electronic characterization studies by high-angle annular dark-field scanning transmission electron microscopy, X-ray photoelectron spectroscopy and X-ray absorption spectroscopy reveal strong Pt–MoO3 interactions, characterized by partial electron transfer. Catalytic tests reveal a volcano-type dependence on the Pt/Mo ratio, with the 4-Pt–MoO3 catalyst achieving the highest performance. The experiments of H2 temperature programmed desorption and in situ diffuse reflectance infrared Fourier transform spectroscopy combined with theoretical calculations support a bifunctional mechanism, in which Pt serves as the primary site for H2 activation, while MoO3 facilitates adsorption and stabilization of polar alanine. Further tuning via thermal treatments shows that the moderate treatment at 500 °C optimally balances the redox state of MoO3 without compromising Pt dispersion, leading to enhanced hydrogenation performance. This work not only advances understanding of metal–oxide interfacial catalysis but also provides a rational design strategy for efficient and selective hydrogenation of amino acids.


1 Introduction

Transformation of amino acids into value-added amino alcohols is an important process for the synthesis of chiral intermediates, pharmaceutical building blocks and nitrogen-containing fine chemicals.1–4 However, conventional organic synthesis of amino alcohols from amino acids remains challenging due to the strong polarity of carboxyl and amino functional groups, which often results in the formation of undesired products from side reactions such as deamination or over-reduction under harsh conditions.5–9 To overcome these limitations, the development of high-performance catalysts that enable selective hydrogenation under mild conditions is highly desirable.10–12 Although catalytic hydrogenation offers a more efficient and sustainable alternative to conventional organic synthesis, such a process normally requires elevated hydrogen pressures (>6 MPa), under which by-products such as amines are often readily formed alongside the desired amino alcohols.13,14 Recently, bifunctional catalysts composed of both metal and metal oxide components, which are capable of coupling hydrogen activation with substrate adsorption, have emerged as a promising strategy to achieve efficient and selective hydrogenation under mild reaction conditions.15–18 Building on these pioneering studies in the hydrogenation of other substrates, to design a bifunctional catalytic system would be promising for enhancing the catalytic performance for the hydrogenation of amino acids.

Among metal–oxide bifunctional systems, the metal sites, such as Pd and Pt sites, typically facilitate the dissociation and activation of hydrogen molecules, while the metal oxide components contribute to the adsorption and activation of polar substrates, such as carbonyl-, carboxyl- and nitro-containing molecules.19,20 A representative example is the Pt1Al1/MgO catalyst, in which highly dispersed Pt sites confined on the MgO(100) surface facilely activate H2 to generate reactive hydrogen species. Such species then migrate to the Al–O–Pt moieties on MgO to hydrogenate the polar substrates adsorbed there, demonstrating a clear synergy between metal sites and oxide sites for catalytic hydrogenation of 3-nitrostyrene.16,21 Another example is the Pt/CeO2 catalyst, where Pt clusters activate H2 to form reactive hydrogen species that spillover to hydrogenate nitro groups of the substrate bound to CeO2 vacancies, also highlighting the cooperative roles of metal and metal oxide in the hydrogenation of the substrate.15 Similarly, on the Pt–MoOx bifunctional catalyst, Pt supplies activated H2, while Mo vacancy sites bind and polarize the acyl carbonyl, enabling a reverse Mars–van Krevelen hydrodeoxygenation (HDO) pathway toward ethers. Consistent with such a bifunctional paradigm, the Pt–Mo/ZrO2 system achieves high ether selectivity under remarkably mild conditions,22 and a complementary sulfoxide HDO study likewise evidences this Pt–MoOx cooperation.23 In spite of these bifunctional systems well designed for such successful cases, mechanistic understanding of the hydrogenation of amino acids, particularly with respect to the nature and function of active sites, has not been clearly revealed but remains crucial for the rational design of high-performance catalysts.24

By choosing Pt and Mo oxides as the corresponding metal and metal oxides, we hereby design a series of Pt–MoO3/TiO2 catalysts with varied Pt/Mo ratios supported on rutile TiO2 to explore the synergistic roles of metal–oxide sites in the hydrogenation of amino acids exemplified with L-alanine. The morphologies and detailed structures of the catalysts were characterized by multiple techniques, such as high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM), X-ray photoelectron spectroscopy (XPS) and X-ray absorption spectroscopy analysis. The results reveal the occurrence of electron transfer from Pt to MoO3, which promotes the formation of MoOx species composed of lower-valence Mo species and oxygen vacancies. Hydrogen temperature-programmed desorption (H2-TPD) and in situ diffuse reflectance infrared Fourier transform spectroscopy (in situ DRIFTS) studies demonstrate that the Pt sites dominate the dissociative activation of hydrogen while the partially reduced MoO3 acts as the anchoring site for carboxyl groups of L-alanine (Ala). Furthermore, density functional theory (DFT) calculations were also employed to rationalize such synergy between Pt and MoO3 species for the hydrogenation of Ala. We finally demonstrate that both the compositional balance and thermal pretreatment govern the synergy of Pt and MoO3, with optimal hydrogenation activity achieved at a specific Pt/Mo ratio and reduction temperature. This work provides fundamental insight into interface-driven catalysis and establishes a generalizable strategy for the rational design of selective hydrogenation catalysts for amino acids and even other polar substrates.

2 Experiments

Synthesis of Pt–MoO3/TiO2

The Pt–MoO3/TiO2 catalysts with varying Pt/Mo atomic ratios were synthesized via a sequential incipient wetness impregnation method. Initially, Pt/TiO2 was prepared by impregnating commercial titanium dioxide (TiO2, >99%, P42/mnm, Aladdin, China) with an aqueous solution of chloroplatinic acid (H2PtCl6·6H2O, >99.5%, Macklin, China), followed by aging at room temperature for 12 h, drying at 80 °C overnight, and calcination at 400 °C for 4 h in air. Then, ammonium molybdate ([(NH4)6Mo7O24·4H2O], >99.5%, Macklin, China) was employed as the molybdenum precursor to introduce Mo onto the pre-synthesized Pt/TiO2 by a second impregnation step. By adjusting the concentration of the molybdate solution, Pt/Mo atomic ratios of 6[thin space (1/6-em)]:[thin space (1/6-em)]1, 4[thin space (1/6-em)]:[thin space (1/6-em)]1, 2[thin space (1/6-em)]:[thin space (1/6-em)]1 and 1[thin space (1/6-em)]:[thin space (1/6-em)]1 were obtained. The impregnated samples were aged, dried at 80 °C, ground, and calcined at 400 °C for 4 h to afford the final Pt–MoO3/TiO2 catalysts. All catalysts were prepared with a fixed Pt loading of 5 wt%. The resulting catalysts were denoted as 6-Pt–MoO3, 4-Pt–MoO3, 2-Pt–MoO3, and 1-Pt–MoO3, respectively, where the leading numbers indicate the Pt/Mo atomic ratio.

Catalytic performance evaluation

The catalytic activity was evaluated via the hydrogenation of Ala to AlaOH, which was carried out in a stainless-steel autoclave equipped with a 100 mL polytetrafluoroethylene liner. Typically, 30 mL of Ala aqueous solution was transferred into the reactor liner, followed by the addition of 100 mg catalyst and 450 μL 85 wt% phosphoric acid. After sealing, the autoclave was purged several times with high-purity argon to remove residual air, and then heated to 95 °C. Unless otherwise noted, a small aliquot (450 μL) of 85 wt% H3PO4 was added to protonate Ala,25 to stabilize the amino acid form,26 and to provide a mild, reproducible proton environment that accelerates hydrogenation kinetics. It should be noted that phosphate species are known to bind strongly to TiO2 and can become irreversibly grafted upon drying/calcination.27 However, under these liquid-phase reaction conditions at 95 °C, no dehydration or calcination pretreatment is involved. Furthermore, in situ AcOH-DRIFTS and catalytic tests indicated that TiO2 is not the primary active site for the hydrogenation and the introduced MoOx provides the principal –COOH adsorption/activation sites. Therefore, it is reasonable to believe that any transient adsorption of phosphate on TiO2, if present under our mild liquid-phase conditions, would have limited impact on the catalytic performance of Pt–MoO3/TiO2 bifunctional catalysts. Once the temperature stabilized, the reactor was purged and charged with high-purity hydrogen (>99.999%) to 4 MPa, and stirring was initiated at 800 rpm to commence the reaction. Upon completion, 1 mL of the reaction mixture was collected, filtered through a 0.22 μm syringe filter, and analyzed using ultra-performance liquid chromatography (UPLC, Waters Acquity). A ZORBAX SB-C18 column (4.6 mm × 150 mm, 5 μm) coupled with a photodiode array detector was employed. Since both Ala and its hydrogenation product, AlaOH, exhibit weak ultraviolet absorption in the range of 190–400 nm, derivatization with phenyl isothiocyanate (PITC) of the reaction solution was performed prior to analysis to enhance their UV detectability.28

3 Results and discussion

Morphologies and structures of bifunctional Pt–MoO3/TiO2 catalysts

The morphologies and structures of the as-synthesized Pt–MoO3/TiO2 bifunctional catalysts with different Pt/Mo atomic ratios were characterized by multiple techniques. Fig. 1a–d show the HAADF-STEM images of 6-Pt–MoO3, 4-Pt–MoO3, 2-Pt–MoO3, and 1-Pt–MoO3 catalysts, respectively, where the initial numbers denote the Pt/Mo ratio. As clearly seen in these HAADF-STEM images, the supported nanoparticles are well dispersed on TiO2 with no obvious aggregation for all the samples. The particle size distributions were obtained by analyzing over 200 nanoparticles in each sample from the collected HAADF-STEM images. The determined average particle sizes for 6-Pt–MoO3, 4-Pt–MoO3, 2-Pt–MoO3, and 1-Pt–MoO3 catalysts are 2.5 ± 0.2, 2.7 ± 0.3, 2.8 ± 0.3 and 3.0 ± 0.4 nm, respectively, showing similar particle sizes with narrow size distributions. These results demonstrate similar morphologies for these Pt–MoO3/TiO2 catalysts with varied Pt/Mo atomic ratios.
image file: d5sc05792b-f1.tif
Fig. 1 HAADF-STEM images and corresponding particle size distributions of the (a) 6-Pt–MoO3, (b) 4-Pt–MoO3, (c) 2-Pt–MoO3 and (d) 1-Pt–MoO3 catalysts. (e) HAADF-STEM EDS mapping images, (f) TEM image and (g) HRTEM images of the 4-Pt–MoO3 catalyst.

HAADF-STEM EDS mapping analysis was further employed to determine the structures of catalysts. As illustrated in Fig. 1e, HAADF-STEM elemental mapping reveals a highly uniform distribution of Pt, Mo, Ti, and O throughout the 4-Pt–MoO3 catalyst. The clear spatial overlap between Pt and Mo signals indicates a close proximity between these two active species, which is critical for promoting synergistic effects at the nanoscale. Complementary transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) images further support the uniform dispersion of nanoparticles across the support (Fig. 1f), indicating effective control over particle size and distribution. The HRTEM images offer insights into the crystalline nature of the catalyst (Fig. 1g). Lattice fringes with measured spacings of 0.196 nm and 0.226 nm can be assigned to the (200) and (111) planes of metallic Pt, respectively (JCPDS no. 87-0646). Meanwhile, the 0.230 nm of fringe spacing corresponds to the (060) plane of orthorhombic MoO3 (JCPDS no. 01-0706). The coexistence of well-defined lattice fringes from both components within the same field of view further confirms the structural integration of Pt and MoO3 at the nanoscale. The combination of HAADF-STEM and HRTEM analyses not only verifies the compositional uniformity of the catalyst but also unveils the intimate contact that underlies its bifunctional catalytic performance.

To further probe the detailed structures of Pt–MoO3/TiO2 bifunctional catalysts, X-ray absorption spectroscopy measurements were also performed for the selected 4-Pt–MoO3 catalyst, together with Pt foil, monometallic Pt and PtO2 samples as the reference. As illustrated in Fig. 2a, the Fourier-transformed extended X-ray absorption fine structure (FT-EXAFS) spectrum at the Pt L3-edge of the Pt foil shows a dominant peak at around 2.4 Å corresponding to Pt–Pt coordination, while that of the PtO2 sample exhibits a distinct peak at around 1.6 Å corresponding to Pt–O coordination. For Pt/TiO2 and 4-Pt–MoO3 catalysts, the FT-EXAFS spectra show main coordination peaks at around 2.6 Å, indicating the presence of metallic Pt–Pt bonding.29 Notably, a minor shoulder peak is seen near the Pt–Pt coordination peak in the spectrum of the 4-Pt–MoO3 catalyst, pointing to the Pt–Mo coordination in the bifunctional catalyst. This is further supported by the wavelet transformed EXAFS (WT-EXAFS) contour plots in Fig. 2b. The WT contour of PtO2 exhibits a strong intensity at low k values (ca. 3–5 Å−1), which is characteristic of Pt–O scattering, as oxygen is a light element (Z = 8) that contributes primarily to low-k components due to its weak backscattering ability. Such a signal is hardly seen in the contour plots of the Pt foil, Pt/TiO2 and 4-Pt–MoO3 catalysts, indicating the absence of oxidized Pt species. The reduced coordination number and enhanced quantum confinement in small Pt clusters lead to sharpening of the photoelectron wave packet and suppression of low-k scattering paths. More importantly, the scattering center of the contour plot for the 4-Pt–MoO3 catalyst shifts slightly toward lower k values, as compared to those of the Pt foil and monometallic Pt/TiO2 sample. Such a shift could be attributed to the presence of minor Pt–Mo coordination,30–32 which is reasonably caused by the intimate contact of Pt and MoO3 demonstrated by HAADF-STEM and HRTEM images.


image file: d5sc05792b-f2.tif
Fig. 2 (a) Fourier-transformed k3-weighted Pt L3-edge EXAFS spectra in R-space, (b) corresponding WT-EXAFS contour plots, and (c) normalized XANES spectra of the PtO2, Pt foil, Pt/TiO2 and 4-Pt–MoO3 catalysts.

The structural indications of the above-observed Pt–MoO3 interaction are further supported by X-ray absorption near-edge structure (XANES) and XPS tests. The Pt L3-edge XANES spectra in Fig. 2c reveal distinct variations in the electronic structure of Pt across the catalyst series. As expected, the spectrum of the PtO2 reference exhibits a pronounced white-line intensity and a positively shifted absorption edge, characteristic of a high oxidation state close to Pt4+. In contrast, that of the Pt foil shows the lowest white-line intensity and an unshifted edge, consistent with a fully metallic Pt0 state. The spectra of Pt/TiO2 and 4-Pt–MoO3 both present slightly elevated white-line intensities and marginally shifted absorption edges relative to Pt foil, suggesting that Pt in these samples remains predominantly metallic.33 Notably, the incorporation of MoO3 results in a subtle but discernible shift in the absorption edge of 4-Pt–MoO3 toward higher energy compared to Pt/TiO2. This delayed edge position suggests a minor increase in the average oxidation state of Pt, pointing to electron transfer from Pt to adjacent MoO3 species.

The XPS results further substantiate the XANES findings. In the Pt 4f region (Fig. 3a), all catalysts exhibit the characteristic doublet peaks corresponding to Pt0 and Pt2+ species.34–36 The XPS spectrum of Pt/TiO2 displays a dominant Pt0 peak at a binding energy of 70.89 eV, reaffirming the metallic nature of Pt in the absence of MoO3. However, upon MoO3 addition, a progressive positive shift in the Pt 4f7/2 binding energy is observed, from 71.13 eV for 6-Pt–MoO3 to 71.38 eV for 1-Pt–MoO3.


image file: d5sc05792b-f3.tif
Fig. 3 (a) Pt 4f and (b) Mo 3d XPS spectra of the 6-Pt–MoO3, 4-Pt–MoO3, 2-Pt–MoO3 and 1-Pt–MoO3 catalysts.

Concurrently, the contribution of Pt2+ species becomes increasingly pronounced at higher Mo contents (i.e., lower Pt/Mo ratios). In parallel, the Mo 3d XPS spectra provide insights into the evolution of electronic properties of Mo species in the presence of Pt.37–39 All Mo-containing samples predominantly exhibit Mo6+ signals, with minor contributions from Mo5+ due to the partial reduction of MoO3 (Fig. 3b). Interestingly, the Mo 3d XPS spectra of the 4-Pt–MoO3 and 6-Pt–MoO3 catalysts with high Pt content even show a weak signal assigned to the Mo0 component, indicative of complete reduction of minor MoO3 to Mo0 on the surface, which is likely facilitated by the presence of Pt. Such Pt-induced reduction of MoO3 is thought to generate oxygen vacancies on the MoO3 surface, which can act as active sites for reactant adsorption, such as amino acid species.40–42 Furthermore, as the Mo content increases, the intensity of reduced Mo5+ and Mo0 features diminishes, implying that excess MoO3 buffers the extent of Pt-induced reduction, likely due to electron dilution and spatial separation effects at higher Mo loadings. Together, the XANES and XPS results offer compelling evidence of electronic interactions between Pt and MoO3, which is also evidenced by Bader charge analysis. The theoretical results show clear electron depletion on interfacial Pt and accumulation on Mo/O of MoO3 (Fig. S29), indicating the interfacial charge transfer.43,44

Bifunctional Pt–MoO3/TiO2 catalysts for L-alanine hydrogenation

Having identified structures of the bifunctional Pt–MoO3/TiO2 catalysts, we turned our attention to investigate the performance of these catalysts for the hydrogenation of amino acids exemplified with Ala here. As shown in Fig. 4a, the monometallic Pt/TiO2 catalyst exhibits low activity, with an Ala conversion below 10%, despite retaining relatively high selectivity toward AlaOH. Upon incorporation of MoO3, the catalytic activity is markedly enhanced. Among all catalysts, the 4-Pt–MoO3 catalyst demonstrates the highest performance, achieving 53.1% of Ala conversion with 96.3% of AlaOH selectivity. The corresponding formation rate of AlaOH reaches 3.61 molAlaOH per molPt+Mo·per h, representing an order of magnitude improvement over the Pt/TiO2 catalyst (0.63 molAlaOH per molPt+Mo·per h). As the Pt/Mo ratio decreases from 4 to 2 and 1, i.e., the 2-Pt–MoO3 and 1-Pt–MoO3 catalysts, the catalytic activity in terms of both Ala conversion and AlaOH formation rate declines sharply, leading to a volcano-type dependence on the Pt/Mo ratio. Notably, while the catalytic activity decreases sharply at higher MoO3 contents, the selectivity remains consistently high across all catalysts, suggesting that product distribution is largely insensitive to the Pt/Mo ratio.
image file: d5sc05792b-f4.tif
Fig. 4 (a) Catalytic performance of Pt–MoO3/TiO2 catalysts with varying Pt/Mo ratios for Ala hydrogenation: the conversion (purple bars), AlaOH selectivity (blue bars), and reaction rate (red line). (b) Reaction energy profiles for H2 dissociation and adsorption on Pt (blue) and MoO3 (purple) surfaces. (c) H2-TPD profiles of bifunctional Pt–MoO3/TiO2 and Pt/TiO2 catalysts. (d) Calculated adsorption energies of Ala and AlaOH and (e) in situ AcOH-DRIFTS spectra of Pt/TiO2 and 4-Pt–MoO3 catalysts.

While the experimental data clearly suggest that the incorporation of MoO3 boosts catalytic activity without compromising selectivity, the precise roles of Pt and MoO3 in the hydrogenation process remain to be clarified. In particular, whether MoO3 directly contributes to hydrogen activation, or instead plays a supporting role in substrate adsorption or intermediate stabilization, is a key question. To address this, a combination of DFT calculations, H2-TPD, and in situ AcOH-DRIFTS tests was employed. Based on Wulff constructions (Fig. S30, Tables S3 and S4), Pt(111) and MoO3(060) were selected as the most thermodynamically stable and experimentally relevant surfaces for DFT studies. The DFT-calculated reaction energy profiles reveal distinct differences in hydrogen activation across these two components (Fig. 4b). On the Pt(111) surface, H2 dissociation occurs spontaneously, with a substantial exothermic energy change from 0 to −0.99 eV, indicating the facile formation of active hydrogen species. In contrast, the MoOx surface exhibits an energetically unfavorable H2 activation pathway, with a high barrier of 1.31 eV and a slightly endothermic adsorption energy (+0.58 eV), confirming that MoOx is not intrinsically active for H2 activation. These findings indicate that Pt serves as the principal site for hydrogen activation, supplying reactive hydrogen atoms for subsequent hydrogenation steps. H2-TPD was also conducted to investigate hydrogen adsorption behavior. As shown in Fig. 4c, the profile of the pure TiO2 support exhibits a weak and broad desorption feature centered at around 100 °C, which likely originates from weakly physisorbed hydrogen species. In contrast, the Pt/TiO2 catalyst shows a distinct desorption peak centered at 190 °C, which can be assigned to the recombinative desorption of dissociated hydrogen from Pt surface sites. Upon incorporation of MoO3, the desorption temperature systematically shifts toward lower values. Specifically, the profiles of 6-Pt–MoO3 and 4-Pt–MoO3 catalysts exhibit desorption peaks at approximately 170 °C and 154 °C, respectively, indicating progressively weakened hydrogen adsorption strength. For the 2-Pt–MoO3 and 1-Pt–MoO3 catalysts, the desorption features become broader and less defined, and no clear peak maxima are observed, suggesting a substantial suppression of hydrogen uptake. This downward shift in desorption temperature with increasing MoO3 content implies that the interaction between hydrogen and Pt is modulated by the presence of adjacent MoO3 species. A plausible explanation is that electronic perturbation of Pt caused by MoO3 alters the hydrogen binding energy, consistent with the increased Pt oxidation state determined by XPS and XANES tests. Collectively, these results indicate that while Pt remains the primary site for hydrogen activation, excessive MoO3 incorporation adversely affects its function by weakening hydrogen adsorption. This finding correlates well with the observed volcano-type activity trend, highlighting the need for an optimal balance between Pt and MoO3 to achieve efficient hydrogen activation and catalytic performance.

To further elucidate the role of the catalyst in substrate activation and product stabilization, DFT calculations were carried out to evaluate the adsorption energies of Ala and its hydrogenation product AlaOH on Pt and MoOx surfaces. As shown in Fig. 4d, both molecules exhibit significantly stronger adsorption on MoO3 compared to Pt. Specifically, the adsorption energy of Ala on MoOx reaches −2.56 eV, while that on Pt(111) is only −1.87 eV. Similarly, AlaOH adsorbs with −2.88 eV on MoOx versus −1.94 eV on Pt(111). These results suggest that MoO3 provides a more favourable surface for anchoring polar functional groups such as carboxyl (–COOH) and hydroxyl (–OH), likely due to its electron-deficient metal centers and oxophilic nature. In contrast, the relatively weaker adsorption on Pt implies limited interaction with the polar reactants and products. This mechanistic insight aligns well with the experimental observation that monometallic Pt/TiO2 exhibits very low activity (<10% conversion) despite retaining high selectivity, indicating that Pt alone is insufficient to activate the substrate effectively. This preferential adsorption on MoO3 is likely to enhance substrate activation and stabilize key intermediates or products, which in turn contributes to the high selectivity observed in experiments.

In situ DRIFTS was conducted to study the adsorption behaviors of Ala by using acetic acid (AcOH) as a probe because its simple structure enables clear identification of carboxyl vibrational features. Previous studies have validated AcOH as a reliable model to represent the adsorption and activation of the –COOH group in amino acids.13,45 For the Pt/TiO2 catalyst, only weak vibrational bands are observed after exposure to AcOH (Fig. 4e), including a weak C[double bond, length as m-dash]O stretching vibration at 1730 cm−1 (ν(C[double bond, length as m-dash]O)), asymmetric and symmetric carboxylate stretching modes at 1645 and 1510 cm−1 (νas(COO) and νs(COO)), respectively, and a methyl bending mode at 1448 cm−1 (δ(CH3)). Although the characteristic bands of adsorbed AcOH can be identified on the Pt/TiO2 catalyst, their intensities are relatively weak and decay rapidly within 30 min under Ar flow, indicating weak and reversible adsorption. In contrast, the spectra of the 4-Pt–MoO3 catalyst exhibit significantly stronger and more persistent vibrational features. Intense bands at 1645 and 1510 cm−1 corresponding to νas(COO) and νs(COO), respectively, suggest the formation of stable carboxylate species. The C[double bond, length as m-dash]O stretching band at 1730 cm−1 is also more pronounced and remains detectable after prolonged Ar purge, along with a clearly resolved δ(CH3) signal at 1448 cm−1.13,45 These vibrational features reflect the coexistence of both molecular and dissociative adsorption states, in contrast to the weak, transient interactions observed on Pt/TiO2, indicating MoO3 as the anchoring site for polar substrates to stabilize key reaction intermediates.

Mechanistic studies based on DFT calculations, H2-TPD, and in situ AcOH-DRIFTS reveal that Pt acts as the primary site for H2 dissociation, whereas MoO3 preferentially serves as the anchoring site for polar functional groups, such as the carboxyl. The observed volcano-shaped trend between activity and the Pt/Mo ratio reflects the need to balance these two complementary functionalities. At an optimal composition (4-Pt–MoO3), the rates of hydrogen activation and substrate adsorption are well-matched, resulting in the highest activity and conversion. In contrast, insufficient or excessive MoO3 disrupts this balance, either limiting substrate anchoring or suppressing hydrogen activation, thereby reducing overall catalytic efficiency.

Thermal treatment governed Pt–MoO3 synergy for hydrogenation

The 4-Pt–MoO3 catalyst was further reduced at different temperatures ranging from 300 to 600 °C to tune the synergy between Pt and MoO3 for the Ala hydrogenation. The morphological characteristics of the 4-Pt–MoO3 catalyst reduced at different temperatures were examined by HAADF-STEM imaging. The nanoparticles are uniformly dispersed on TiO2 across all the samples, with no signs of significant aggregation, even after high-temperature reduction (Fig. 5a–d). Particle size distributions were obtained by analyzing over 200 individual nanoparticles in each sample. The average particle sizes for the samples reduced at 300, 400, 500 and 600 °C are 2.3 ± 0.2, 2.5 ± 0.3, 2.7 ± 0.4 and 2.7 ± 0.5 nm, respectively, indicating similar morphologies, and therefore, particle size effects can be excluded as the primary factor influencing catalytic behavior. In addition, XRD patterns only show the diffraction peaks of rutile TiO2 (JCPDS no. 01-1292) and Pt (JCPDS no. 04-0802) without those assigned to MoO3 at all reduction temperatures (Fig. S14). This indicates that Mo species remain amorphous or as highly dispersed MoOx, with domain sizes below the detection limit of XRD. HRTEM analyses confirm that reduction at 300–600 °C tunes the oxygen vacancy concentration in MoOx while preserving the crystal structures of TiO2 and Pt (Fig. S5), with MoOx lattice fringes (0.22–0.23 nm reported previously) consistently observed.46 The influence of thermal reduction on the electronic structure of the catalyst was then evaluated by XPS analysis. As shown in Fig. 5e, the Pt 4f XPS spectra exhibit only slight changes with increasing temperature from 300 to 600 °C. All samples display characteristic Pt0 and Pt2+ doublet peaks, with the Pt0 4f7/2 component located near 71.13 eV. The minor changes in binding energy and relative peak intensity suggest that the oxidation state and electronic environment of Pt remain largely unaffected under the current conditions. Unlike the strong electronic modulation induced by increasing MoO3 content (Fig. 3a), the reduction temperature appears insufficient to significantly alter the electron density around Pt sites in the fixed Pt/Mo ratio. In contrast, the Mo 3d XPS spectra reveal a clear evolution in the oxidation state of Mo species with increasing reduction temperature (Fig. 5f). At 300 °C, the spectra are dominated by Mo6+ signals, characteristic of stoichiometric MoO3. As the temperature increases to 400 °C and beyond, additional peaks corresponding to partially reduced Mo5+ and Mo0 components become significant, indicating thermal reduction of MoO3 under a H2 atmosphere.
image file: d5sc05792b-f5.tif
Fig. 5 HAADF-STEM images and corresponding particle size distributions of 4-Pt–MoO3 reduced at (a) 300, (b) 400, (c) 500 and (d) 600 °C. (e) Pt 4f and (f) Mo 3d XPS spectra of 4-Pt–MoO3 catalysts reduced at various temperatures. (g) Catalytic performance of 4-Pt–MoO3 catalysts reduced at various temperatures.

The Mo 3d XPS spectra were deconvoluted to quantify the fractions of Mo6+ and Mo5+ species that varied with the reduction temperature. Clearly, with the increasing H2 reduction temperature, the amount of Mo6+ decreases while that of Mo5+ increases (Fig. 5f). In addition, the average valence of oxidized Mo species (vavg) is defined to compare the differences in the electronic structure caused by various reduction temperatures,47 and the value of vavg decreases with increasing reduction temperatures (Table S1), evidencing the Pt-promoted formation of oxygen vacancies at elevated temperature. This is also confirmed by the EPR tests, which demonstrate that the signal at g = 2.003 appears to be stronger with increasing reduction temperatures (Fig. S27). Such reduction behavior is consistent with previous reports of MoO3 reducibility under moderate conditions.37,38 The emergence of Mo0 signals at high temperatures suggests the formation of oxygen-deficient surface species, potentially associated with the generation of surface oxygen vacancies. These vacancies have been proposed as active sites for polar substrate adsorption and could contribute to the observed catalytic performance. It is also worth noting that, unlike the Mo content-dependent XPS trends shown in Fig. 3b, where excessive loading suppressed Pt-induced reduction via electronic dilution, the thermal reduction effect observed here is more direct and progressive, driven by temperature rather than compositional changes. Together, these results underscore the redox flexibility of Mo species under reductive conditions and highlight the critical role of thermal treatment in tuning surface electronic states and potential adsorption sites.

The influence of reduction temperature on the catalytic performance of the 4-Pt–MoO3 catalyst was further evaluated. As shown in Fig. 5g and S12, thermal treatment at 500 °C affords the best catalytic performance, with a favorable balance among conversion, selectivity and reaction rate. Under this condition, the catalyst achieves a conversion of 53.1% and a selectivity of 96.3%. As the treatment temperature increases beyond 500 °C, the selectivity continues to improve, reaching nearly 100% at 600 °C. However, this comes at the expense of catalytic activity, as evidenced by a marked decline in conversion (to 36.1%) and a prolonged time to reach full conversion (>22 h). This performance drop is likely due to the excessive reduction of MoO3, which may disrupt the electronic interactions between Pt and Mo species or lead to blockage of active adsorption sites on the oxide surface. This trend suggests that moderate treatment effectively balances hydrogen activation on Pt and substrate interaction on MoO3, preserving the bifunctional synergy critical for efficient hydrogenation. In contrast, under- or over-treatment leads to suboptimal performance, either due to insufficient electronic modification of MoO3 or excessive reduction that impairs functionality.

In addition, since the surface TiOx species is possible to be present under high-temperature reduction, it is necessary to address the role of such possibly formed species. Thus, XPS tests for Ti 2p orbitals for the catalysts reduced at various temperatures were further performed. As shown in Fig. S15, the Ti 2p3/2 XPS peak shifts slightly from 458.38 to 458.33 eV with increasing reduction temperature from 300 to 600 °C. This very slight shift indicates that while the Ti oxidation state becomes marginally more positive at higher reduction temperatures, the extent of reduction remains limited. Such very small fraction of surface TiOx species is believed to deliver minor effects on the catalytic performance. The in situ AcOH-DRIFTS spectrum of the Pt–MoO3/TiO2 catalyst exhibits marked vibrational features associated with carboxyl adsorption compared with Pt/TiO2, particularly the ν(C[double bond, length as m-dash]O) stretching and the νas/s(COO) bands (Fig. 4e). Furthermore, the Pt/TiO2 catalyst exhibits only negligible Ala conversion, underscoring that neither metallic Pt sites nor the limited TiOx species generated on TiO2 provide sufficient carboxyl adsorption or activation. Upon introducing MoO3, both the conversion and the intrinsic reaction rate increase markedly (Fig. 4a). These results unveil that MoO3 substantially enhances the adsorption and activation of the carboxyl group, whereas the contribution of TiOx remains relatively limited.

Considering that the presence of water as the solvent can decrease the amount of oxygen vacancies in the catalyst,48 the cycle-tests were further performed for the optimal catalyst (i.e., 4-Pt–MoO3 reduced at 500 °C) to address the catalytic stability. The tests across five consecutive batch runs (4 MPa H2, 95 °C, 8 h) reveal that AlaOH yield declines monotonically from 50% to 21% without H2 regeneration (Fig. S13). XPS tests for the catalyst after the reaction reveal a slight increase in the valency, and the EPR spectra show a marked attenuation of the signal at g = 2.003 assigned to the Mo5+ oxygen vacancy after the reaction (Fig. S16 and S26).49 These results indicate the decrease of oxygen vacancies in the catalyst after the reaction. In contrast, the initial activity could be recovered when a regeneration under hydrogen at 500 °C was performed between each run, which can be explained by the enriched oxygen vacancies by high-temperature reduction. The contrasting stability results also verify that the hydrogenation of L-alanine follows the reverse Mars–van Krevelen mechanism as schematically shown in Fig. S8.50–53

4 Conclusions

In summary, a series of bifunctional Pt–MoO3/TiO2 catalysts were systematically investigated to uncover the structural and electronic origins of their performance in alanine hydrogenation. Advanced characterization techniques, including HAADF-STEM, HRTEM, XPS, and XANES, revealed that the intimate contact between Pt and MoO3 induces electronic interactions, characterized by partial electron transfer from Pt to MoO3. The H2-TPD analysis combined with DFT calculations and in situ AcOH-DRIFTS tests support a bifunctional mechanism, wherein Pt activates hydrogen while MoO3 facilitates substrate adsorption and stabilization. Catalytic testing demonstrated a volcano-type dependence on the Pt/Mo ratio, with the 4-Pt–MoO3 catalyst achieving the highest conversion and selectivity. Further tuning via thermal reduction treatments showed that moderate reduction at 500 °C optimally balances the redox state of MoO3 without compromising Pt dispersion, leading to enhanced hydrogenation performance. In contrast, excessive reduction at higher temperatures disrupted the delicate synergy and diminished catalytic activity. The insights gained here offer valuable guidance for the rational design of metal–oxide bifunctional systems in selective hydrogenation and even other structure-sensitive catalytic transformations.

Author contributions

Conceptualization: XZ, XD, and YC. Supervision: XZ, XD, and YC. Funding acquisition: XZ, XD, and YC. Projection administration: YC. Resources: XZ, XD, and YC. Methodology: YJ, RS, XG, and NA. Investigation: YJ, RS, JM, and NZ. Data curation: YJ and RS. Formal analysis: YJ, RS, and NZ. Software: RS and NZ. Visualization: YJ, RS, and XG. Writing – original draft: YJ and RS. Writing – review & editing: all authors.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this article have been included as part of the supplementary information (SI). Supplementary information: detailed experimental procedures, catalyst synthesis and calculation model, additional characterisation (XRD, XPS, EPR, TEM), supplementary figures and tables underlying the main results. See DOI: https://doi.org/10.1039/d5sc05792b.

Acknowledgements

This work was financially supported by the National Key R&D Program of China (2022YFA1503502), the Natural Science Foundation of China (22332003, 22478106, 22538004 and 22408097), the Chang Jiang Scholars Program of the Ministry of Education of China (T2022163), Shanghai Rising-star Program (23QA1401900), Young Elite Scientists Sponsorship Program by CAST (2023QNRC001), the Open Project of Yunnan Precious Metals Laboratory Co., Ltd (YPML-2023050272), and Foundation of Yunnan Province Science and Technology Department (202305AF150194). We thank the BL11B and BL14W1 XAFS beamlines of Shanghai Synchrotron Radiation Facility (SSRF) for providing the beamtime. We thank the Shanghai Synchrotron Radiation Facility (https://cstr.cn/31124.02.SSRF.BL16U1) for assistance in EXAFS measurements.

Notes and references

  1. H. A. Rudbari, N. Kordestani, J. V. Cuevas-Vicario, M. Zhou, T. Efferth, I. Correia, T. Schirmeister, F. Barthels, M. Enamullah, A. R. Fernandes and N. Micale, Investigation of the influence of chirality and halogen atoms on the anticancer activity of enantiopure palladium(ii) complexes derived from chiral amino-alcohol Schiff bases and 2-picolylamine, New J. Chem., 2022, 46, 6470–6483 RSC.
  2. S. Hameury, H. Bensalem and K. De Oliveira Vigier, Sustainable Amination of Bio-Based Alcohols by Hydrogen Borrowing Catalysis, Catalysts, 2022, 12, 1306 CrossRef CAS.
  3. F. Kallmeier, B. Dudziec, T. Irrgang and R. Kempe, Manganese-Catalyzed Sustainable Synthesis of Pyrroles from Alcohols and Amino Alcohols, Angew. Chem., Int. Ed., 2017, 56, 7261–7265 CrossRef CAS PubMed.
  4. C. Nájera and J. M. Sansano, Catalytic Asymmetric Synthesis of α-Amino Acids, Chem. Rev., 2007, 107, 4584–4671 CrossRef PubMed.
  5. J. Wang, X. Lin, P.-L. Shao, J. Song, J. Wen and X. Zhang, Double Asymmetric Hydrogenation of α-Iminoketones: Facile Synthesis of Enantiopure Vicinal Amino Alcohols, ACS Catal., 2021, 11, 12729–12735 CrossRef CAS.
  6. D. Peña, A. J. Minnaard, J. G. de Vries and B. L. Feringa, Highly Enantioselective Rhodium-Catalyzed Hydrogenation of β-Dehydroamino Acid Derivatives Using Monodentate Phosphoramidites, J. Am. Chem. Soc., 2002, 124, 14552–14553 CrossRef PubMed.
  7. M. J. Burk, J. G. Allen and W. F. Kiesman, Highly Regio- and Enantioselective Catalytic Hydrogenation of Enamides in Conjugated Diene Systems: Synthesis and Application of γ,δ-Unsaturated Amino Acids, J. Am. Chem. Soc., 1998, 120, 657–663 CrossRef CAS.
  8. M. J. Burk, J. E. Feaster, W. A. Nugent and R. L. Harlow, Preparation and use of C2-symmetric bis(phospholanes): production of .alpha.-amino acid derivatives via highly enantioselective hydrogenation reactions, J. Am. Chem. Soc., 1993, 115, 10125–10138 CrossRef CAS.
  9. M. D. Fryzuk and B. Bosnich, Asymmetric synthesis. Production of optically active amino acids by catalytic hydrogenation, J. Am. Chem. Soc., 1977, 99, 6262–6267 CrossRef CAS PubMed.
  10. J. Verduyckt, R. Coeck and D. E. De Vos, Ru-Catalyzed Hydrogenation–Decarbonylation of Amino Acids to Bio-based Primary Amines, ACS Sustain. Chem. Eng., 2017, 5, 3290–3295 CrossRef CAS.
  11. K. P. Pimparkar, D. J. Miller and J. E. Jackson, Hydrogenation of Amino Acid Mixtures to Amino Alcohols, Ind. Eng. Chem. Res., 2008, 47, 7648–7653 CrossRef CAS.
  12. A. Vandekerkhove, L. Claes, F. De Schouwer, C. Van Goethem, I. F. J. Vankelecom, B. Lagrain and D. E. De Vos, Rh-Catalyzed Hydrogenation of Amino Acids to Biobased Amino Alcohols: Tackling Challenging Substrates and Application to Protein Hydrolysates, ACS Sustain. Chem. Eng., 2018, 6, 9218–9228 CrossRef CAS.
  13. M. Tamura, R. Tamura, Y. Takeda, Y. Nakagawa and K. Tomishige, Insight into the mechanism of hydrogenation of amino acids to amino alcohols catalyzed by a heterogeneous MoO(x) -modified Rh catalyst, Chemistry, 2015, 21, 3097–3107 CrossRef CAS PubMed.
  14. M. Tamura, R. Tamura, Y. Takeda, Y. Nakagawa and K. Tomishige, Catalytic hydrogenation of amino acids to amino alcohols with complete retention of configuration, Chem. Commun., 2014, 50, 6656–6659 RSC.
  15. X. Ren, J. Huang, J. Ma, Y. Zhang, W. Chu, S. Perathoner, G. Centi and Y. Liu, Boosting the activity in the liquid-phase hydrogenation of S-containing nitroarenes by dual-site Pt/CeO2 catalysts design, Nat. Commun., 2025, 16, 4851 CrossRef CAS PubMed.
  16. D. Gao, S. Wang, Y. Lv, C. Wang, J. Ren, P. Zheng, L. Song, A. Duan, X. Wang, G. Chen and C. Xu, Unraveling Hydrogenation Kinetic Behavior of Transition Metal Oxides via Decoupling Dihydrogen Dissociation and Substrate Activation, ACS Catal., 2024, 14, 1939–1950 CrossRef CAS.
  17. H. Choi, Y. Choi, J. Kim, J. H. Lee, E. Kang, J. Yun, H. Park, M. Kim, H. Ullah, K. Shin and H. Y. Kim, Mechanistic insight into bifunctional thermocatalytic CO2 hydrogenation by Oxide-Supported Palladium, Chem. Eng. J., 2025, 503, 158163 CrossRef CAS.
  18. S. Ebrahimian, J. Vongsvivut, T. Williams and S. Bhattacharya, High-Pressure and Low-Temperature CO2 Hydrogenation to DME over an Oxalate-Derived Bifunctional Catalyst: In Situ Synchrotron-FTIR Analysis, Energy Fuels, 2025, 39(28), 13688 CAS.
  19. A. W. Savoy, Fundamental Catalytic Properties of Faceted Anatase Nanoparticles and Their Influence on Supported Molybdena, PhD thesis, Washington State University, 2023.
  20. M. İ. Şener, Development of Bifunctional Catalyst for the Single-step Synthesis of Dimethyl Ether, MS thesis, Middle East Technical University, Turkey, 2019.
  21. R. Xue, F.-F. Wang, D.-L. Chen, F. Zhang and W. Zhu, Theoretical Insights on the Synergistic Effect of Dual Metal Sites Supported on MgO(100) Promoting the Hydrogenation Reaction, J. Phys. Chem. C, 2024, 129, 359–368 CrossRef.
  22. K. Sakoda, S. Yamaguchi, T. Mitsudome and T. Mizugaki, Selective Hydrodeoxygenation of Esters to Unsymmetrical Ethers over a Zirconium Oxide-Supported Pt-Mo Catalyst, JACS Au, 2022, 2, 665–672 CrossRef CAS PubMed.
  23. Y. Kuwahara, Y. Yoshimura, K. Haematsu and H. Yamashita, Mild Deoxygenation of Sulfoxides over Plasmonic Molybdenum Oxide Hybrid with Dramatic Activity Enhancement under Visible Light, J. Am. Chem. Soc., 2018, 140, 9203–9210 CrossRef CAS PubMed.
  24. A. S. Y. Yim and M. Wills, Asymmetric transfer hydrogenation using amino acid derivatives; further studies and a mechanistic proposal, Tetrahedron, 2005, 61, 7994–8004 CrossRef CAS.
  25. F. T. Jere, D. J. Miller and J. E. Jackson, Stereoretentive C− H bond activation in the aqueous phase catalytic hydrogenation of amino acids to amino alcohols, Org. Lett., 2003, 5, 527–530 CrossRef CAS PubMed.
  26. J. E. Holladay, T. A. Werpy and D. S. Muzatko, Catalytic hydrogenation of glutamic acid, Appl. Biochem. Biotechnol., 2004, 115, 857–869 CrossRef PubMed.
  27. Z. Wang, A. Mahmood, X. Xie, X. Wang, H. Qiu and J. Sun, Surface adsorption configurations of H3PO4 modified TiO2 and its influence on the photodegradation intermediates of gaseous o-xylene, Chem. Eng. J., 2020, 393, 124723 CrossRef CAS.
  28. R. Song, C. Yao, W. Li, N. An, Y. Shen, N. Fei, X. Ge, Y. Cao, X. Duan and X. Zhou, Kinetic insights into structure sensitivity of Ru catalyzed l-alanine hydrogenation to alaninol, React. Chem. Eng., 2025, 10, 135–145 RSC.
  29. Y. Li, Y. Cao, X. Ge, H. Zhang, K. Yan, J. Zhang, G. Qian, Z. Jiang, X. Gong, A. Li, X. Zhou, W. Yuan and X. Duan, Pt-O4 moiety induced electron localization toward In2O-Triggered acetylene Semi-Hydrogenation, J. Catal., 2022, 407, 290–299 CrossRef CAS.
  30. R. Li, J. Shang, F. Wang, Q. Lu, H. Yan, Y. Tuo, Y. Liu, X. Feng, X. Chen, D. Chen and C. Yang, Quantification and optimization of platinum-molybdenum carbide interfacial sites to enhance low-temperature water-gas shift reaction, Nat. Commun., 2025, 16, 1098 CrossRef CAS PubMed.
  31. H. Lin, P. Sun, Y. Xu, X. Zong, H. Yang, X. Liu, H. Zhao, L. Tan, L. Wu and Y. Tang, Enhanced selective cleavage of aryl C-O bond by atomically dispersed Pt on α-MoC for hydrodeoxygenation of anisole, Mol. Catal., 2022, 531, 112652 CAS.
  32. M. Ma, G. Li, W. Yan, Z. Wu, Z. Zheng, X. Zhang, Q. Wang, G. Du, D. Liu, Z. Xie, Q. Kuang and L. Zheng, Single-Atom Molybdenum Engineered Platinum Nanocatalyst for Boosted Alkaline Hydrogen Oxidation, Adv. Energy Mater., 2022, 12, 2103336 CrossRef CAS.
  33. W. Chen, J. Cao, J. Yang, Y. Cao, H. Zhang, Z. Jiang, J. Zhang, G. Qian, X. Zhou, D. Chen, W. Yuan and X. Duan, Molecular-level insights into the electronic effects in platinum-catalyzed carbon monoxide oxidation, Nat. Commun., 2021, 12, 6888 CrossRef CAS PubMed.
  34. E. I. Vovk, A. V. Kalinkin, M. Y. Smirnov, I. O. Klembovskii and V. I. Bukhtiyarov, XPS Study of Stability and Reactivity of Oxidized Pt Nanoparticles Supported on TiO2, J. Phys. Chem. C, 2017, 121, 17297–17304 CrossRef CAS.
  35. M.-J. Choi, H. Park, M. H. Engelhard, D. Li, P. V. Sushko and Y. Du, Reevaluation of XPS Pt 4f peak fitting: Ti 3s plasmon peak interference and Pt metallic peak asymmetry in Pt@TiO2 system, J. Vac. Sci. Technol., A, 2024, 42, 063209 CrossRef CAS.
  36. C. Hao, J. Gan, Y. Cao, W. Luo, W. Chen, G. Qian, X. Zhou and X. Duan, Crucial size effects of atomic-layer-deposited Pt catalysts on methanol electrooxidation, Catal. Today, 2021, 364, 157–163 CrossRef CAS.
  37. W. Gruenert, A. Y. Stakheev, R. Feldhaus, K. Anders, E. S. Shpiro and K. M. Minachev, Analysis of molybdenum (3d) XPS spectra of supported molybdenum catalysts: an alternative approach, J. Phys. Chem., 1991, 95, 1323–1328 CrossRef CAS.
  38. J.-G. Choi and L. Thompson, XPS study of as-prepared and reduced molybdenum oxides, Appl. Surf. Sci., 1996, 93, 143–149 CrossRef CAS.
  39. J. Fang, W. Sun, K. Huang, X. Ge, W. Chen, G. Qian, Y. Cao, J. Zeng, J. Ma, X. Zhou, X. Duan, J. Zhang and L. Jiang, Structural and Kinetic Insights into Interfacial Site Effects in NiMo-Catalyzed Hydrodeoxygenation of Palmitic Acid, Ind. Eng. Chem. Res., 2025, 64, 5962–5972 CrossRef CAS.
  40. L. A. Gomez, R. Bababrik, M. R. Komarneni, J. Marlowe, T. Salavati-fard, A. D. D'Amico, B. Wang, P. Christopher and S. P. Crossley, Selective Reduction of Carboxylic Acids to Aldehydes with Promoted MoO3 Catalysts, ACS Catal., 2022, 12, 6313–6324 CrossRef CAS.
  41. R. Liu, X. Li, K. Ma, Z. Chen and C. Tang, Sustainable production of bio-propionic acid: synergy between vacancy and thermoelectron in MoS2/MoO3 composite-enhanced hydrodeoxygenation of lactic acid, Catal. Sci. Technol., 2023, 13, 1221–1232 RSC.
  42. S. Zhong, C. Xing, A. Cao, P. Zeng, W. Cai and Y. Li, A sensitive colorimetric chiral recognition for thiol-containing amino acids based on NIR plasmonic MoO3−x nanoparticles, J. Mater. Chem. C, 2021, 9, 11091–11097 RSC.
  43. K. B. Wiberg, C. M. Hadad, T. J. LePage, C. M. Breneman and M. J. Frisch, Analysis of the effect of electron correlation on charge density distributions, J. Phys. Chem., 1992, 96, 671–679 CrossRef CAS.
  44. H. J. Freund and G. Pacchioni, Oxide ultra-thin films on metals: new materials for the design of supported metal catalysts, Chem. Soc. Rev., 2008, 37, 2224–2242 RSC.
  45. C. Kaku, S. Suganuma, K. Nakajima, E. Tsuji and N. Katada, Selective Hydrogenation of L-proline to L-prolinol over Al2O3-supported Pt-MoOx Catalyst, ChemCatChem, 2022, 14, e202200399 CrossRef CAS.
  46. F. Niu, G. Xu, H. Yang, Y. Liang, Z. Xu, H. Li, X. Wei and L. Yang, Alkylamine-tuned MoOx with synergistic manipulation of interlayer spacing and oxygen vacancies toward advanced Li–S batteries, Appl. Phys. Lett., 2025, 126, 033902 CrossRef CAS.
  47. F. Kröger and H. Vink, in Solid state physics, Elsevier, 1956, vol. 3, pp. 307–435 Search PubMed.
  48. T. Prasomsri, T. Nimmanwudipong and Y. Román-Leshkov, Effective hydrodeoxygenation of biomass-derived oxygenates into unsaturated hydrocarbons by MoO3 using low H2 pressures, Energy Environ. Sci., 2013, 6, 1732 RSC.
  49. A. Adamski and Z. Sojka, EPR studies on NO interaction with MoOx/t-ZrO2 catalysts obtained by slurry deposition, Catal. Today, 2008, 137, 283–287 CrossRef CAS.
  50. Y. Kuwahara, T. Mihogi, K. Hamahara, K. Kusu, H. Kobayashi and H. Yamashita, A quasi-stable molybdenum sub-oxide with abundant oxygen vacancies that promotes CO2 hydrogenation to methanol, Chem. Sci., 2021, 12, 9902–9915 RSC.
  51. P. Mars and D. W. Van Krevelen, Oxidations carried out by means of vanadium oxide catalysts, Chem. Eng. Sci., 1954, 3, 41–59 CrossRef CAS.
  52. S. Mine, T. Yamaguchi, K. W. Ting, Z. Maeno, S. H. Siddiki, K. Oshima, S. Satokawa, K.-i. Shimizu and T. Toyao, Reverse water-gas shift reaction over Pt/MoOx/TiO2: reverse Mars–van Krevelen mechanism via redox of supported MoOx, Catal. Sci. Technol., 2021, 11, 4172–4180 RSC.
  53. A. V. Mironenko and D. G. Vlachos, Conjugation-Driven "Reverse Mars-van Krevelen"-Type Radical Mechanism for Low-Temperature C-O Bond Activation, J. Am. Chem. Soc., 2016, 138, 8104–8113 CrossRef CAS PubMed.

Footnote

These authors contributed equally to this work.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.