Rational engineering of hydroxyapatites for sustainable chemicals, H2, biofuels and CO2 conversion
Abstract
Chemical and fuel production through conventional catalytic processes requires significant improvement to reduce CO2 emission levels. Sustainable, potentially direct, chemical production is in critical need, such as methane coupling to olefins, propane dehydrogenation to propylene, biodiesel production, and greenhouse gas emissions modulating reactions, including CO2-utilizing dry reforming of methane, partial oxidative of methane, CO oxidation, CO2 methanation are some of the emerging technologies to address sustainability challenges. These technologies remained constrained due to the lack of stable and efficient catalysts. Hydroxyapatite (HAP), a highly functional versatile material, offers great potential due to its flexible tunability, and several studies have outlined the synthesis protocol of HAP and design modifications for catalytic applications. However, a comprehensive understanding of connecting reaction-specific demands to tailor HAP catalyst designs is limited. In this review, we bridge that gap, highlight key challenges in catalytic reactions, and propose the necessary HAP catalyst modifications, including acid-base tuning, defects-induced lattice oxygen or vacancies, mesoporosity modulation, and catalyst active metal species dispersion, to improve catalytic performance by limiting catalyst deactivation from absorbates surface poisoning, sintering, and coking. Finally, future research areas for improvement for HAP catalysts are suggested to advance the maturity of catalytic technologies.
Keywords
INTRODUCTION
Attenuating CO2 concentration is crucial for mitigating climate change, the key challenge of the 21st century. Creating sustainable, long-term, and economically viable methods for CO2 capture and valorization is crucial, with heterogeneous catalysts becoming increasingly vital to contemporary sustainable energy strategies[1]. The application of heterogeneous catalysts enables fine chemical and energy production to occur more sustainably. The development of heterogeneous catalysts requires an intimate and fundamental understanding of the catalyst surface at the nano-scale to drive desired catalytic reactions selectively and actively with stable performance. A significant amount of literature has been dedicated to materials for heterogeneous catalysts such as carbons[2,3], metal-organic framework (MOF)[4-6], and metal oxides[3,7,8]. Other than these classes of catalyst support, hydroxyapatite (HAP)-based catalysts have received increasing research attention in the past few decades, as evidenced by an increasing number of publications in Figure 1A.
Figure 1. (A) Number of publications using keywords “hydroxyapatite” and “catalyst” from the year 2000 to the year 2025 from Web of Science as of 28 March 2025. (B) Schematic of the various surface engineering strategies for HAP.
HAP is a bio-compatible material that constitutes the main mineral component in bone and teeth and is responsible for hardness and strength. This bio-compatible material stirs great interest in many fields due to its remarkable structure and inherent properties. The surface of HAP material contains both acidic and basic elements such as P and Ca, respectively[9]. HAP has often been used as a catalyst or catalyst support for liquid-phase organic reactions, namely, the Suzuki reaction, alcohol oxidation and furfural amination reactions[10,11]. Surface engineering strategies, based on a rational understanding of HAP material properties, are required to unlock the potential of HAP materials for enhanced catalytic performance, product selectivity, and catalytic stability. This is particularly important when HAP materials are to be used as heterogeneous catalysts. Ibrahim et al. had written a comprehensive review on the usage of HAP for air and water pollution control[12]. Kaneda and Mizugaki comprehensively covered HAP as a heterogeneous catalyst for organic synthesis reactions and pointed out the possibility of using HAP to solve environmental and energy issues[13]. The synthesis route of HAP material for liquid phase reaction has been critically covered by Fihri et al.[14]
FUNDAMENTAL UNDERSTANDING OF HAP FRAMEWORK
HAP structure
HAP has a hexagonal (P63/m space group) crystal cell with lattice parameters a = 9.37 Å and c = 6.88 Å[17]. In the HAP crystal structure, two distinct types of calcium (Ca) atoms occupy different positions within the unit cell [Figure 2A][18]. Consequently, the chemical formula is expressed as Ca(I)Ca(II)6(PO4)6(OH)2. The HAP framework is composed of a dense network of PO4 tetrahedra, with each PO4 group contributing to a column structure that defines two types of discrete channels. The first channel has a diameter of 2.5 Å with adjacent Ca2+ ions. The adjacent Ca2+ ions are in the Ca(II) location and these Ca2+ play a significant role in modulating the acid-base and electrical properties of HAP material. HAP allows large variations in composition and can be a highly non-stoichiometric solid. Stoichiometric HAP has the chemical formula of Ca10(PO4)6(OH)2 where the ratio Ca/P is 1.67. Non-stoichiometric HAP has Ca/P < 1.67 with chemical formula of Ca10-x(HPO4)x(PO4)6-x(OH)2-x where 0 < x < 1. The OH- groups are arranged along the c-axis in the HAP unit cell to balance the positive charge of the matrix. The substitution of cations or anions within the lattice can lead to vacancy defects at calcium or OH- sites. Alternatively, such substitutions may induce structural disorder, resulting in the formation of an amorphous phase.
Figure 2. (A) HAP hexagonal crystal structure unit cell with two Ca sites, Ca(I) and Ca(II) sites. Reproduced from Ref.[18] under CC BY 4.0 license. (B) Product selectivity at 50% ethanol conversion: (▲) total Guerbet alcohols, (●) 1,3-butadiene (■), ethylene. Reproduced from Ref.[19] with permission from Elsevier. (C) Moles of Ca2+ released to the solution as a function of moles of Pb2+. Reproduced from Ref.[20] with permission from Elsevier. (D) 43Ca MAS spectra of PbxCa10-x(VO4)(PO4)5(OH)2 (x = 0, 2, 4). Reproduced from Ref.[21] with permission from American Chemical Society.
The stoichiometric HAP crystals can exist in two forms: hexagonal (lattice parameters a = b = 9.432 Å,
Importance of the Ca/P ratio
As the apatite material consists of calcium cation and phosphate anion in the unit cell, the surface properties of the apatite material are intimately connected to the apatite surface composition. When the Ca/P is at a stoichiometric ratio of 1.67, there are more basic sites than acidic sites, which results in HAP displaying basic catalyst behavior. HAP surface displays acidic behavior when the Ca/P = 1.50, which constitutes a non-stoichiometric HAP. The mix of basic and acid sites imbues the catalyst surface with amphoteric properties if the Ca/P is between 1.5-1.67. However, it is also important to note that the bulk Ca/P ratio is observed to be greater than the surface Ca/P ratio[19]. Thus, the HAP surface can be non-stoichiometric in nature. HAP crystal tends to grow along the c-axis (i.e., rod-like HAP) and expose more of the a-faces, which are rich in Bronsted acid phosphate groups[19].
The implications of the Ca/P ratio on the surface properties of the HAP support material can be observed from the differences in catalytic activity and product selectivity as noted by several authors[19,29,30]. HAP had been used as a catalyst for Guerbet alcohol synthesis, wherein the presence of more acidic sites in HAP catalyst at varying Ca/P has shown a clear influence in favoring ethylene selectivity while more basic sites favor 1-butanol selectivity as shown in Figure 2B[19]. If an active transition metal such as nickel is incorporated into the HAP support, the Ca/P can also be varied to induce CaO to exsolve from the HAP surface which can help to tune metal-support interaction effects[31,32]. Furthermore, the decrease of Ca/P can induce the presence of more acidic sites which can help to anchor the nickel nanoparticles onto the HAP support, thus improving catalyst stability.
Cationic substitutability
The calcium cation site in the apatite unit cell is flexible enough to accommodate transition metal cations that are similar to its ionic radii. The calcium cations in Ca(I)Ca(II)6(PO4)6(OH)2 HAP are situated in two different places in the framework, Ca(I) and Ca(II). The coordinating number of the Ca(I) is 9, while the one for Ca(II) is 7. Such a difference suggests that Ca(I) would be preferentially occupied by other cations larger than Ca2+[33]. Transition metal cations can be incorporated into the HAP framework using coprecipitation or ion-exchange methods. In coprecipitation, two aqueous solutions are typically prepared, one with Ca2+ and cation, and another with the phosphate precursor. The pH of both solutions should be similar for concurrent precipitation of Ca2+ and P.
The calcium sites in the HAP lattice are also amenable to substitution by another cation from transition metal elements or lanthanide elements[34]. The calcium sites can be substituted by monovalent (e.g., Na+, K+, Ag+, etc.), divalent (e.g., Cu2+, Ni2+, Zn2+, etc.), and trivalent (e.g., Ce3+, Co3+, etc.) metals. However, the extent of difficulty in substituting the Ca sites could be dependent on the electronegativity of the element, assuming the same ionic radius[35]. For monovalent cations such as Na+, K+, and Li+, Na+ has a better affinity for cation substitution into the HAP lattice due to similar ionic radii as Ca2+ (0.9 Å). Thus, the order of affinity decreases in the following manner: Na+ (1.0 Å) > K+ (1.4 Å)> Li+ (0.7 Å)[36]. The preference for cation exchanges due to differences in ionic radii relative to the Ca2+ ionic radii is similar for divalent components. Takeuchi et al. observed that Pb2+ was exchanged faster than Cu2+ and Cd2+ due to the same ionic radii as Ca2+ in the single-component adsorption study[37]. In a dual-component adsorption study, the Cu2+ was more difficult to exchange with Ca2+ and was attributed to the bigger ionic radii of Cu2+ (1.3 Å) as compared to Ca2+ (1.1 Å). However, Cd2+ (1.1 Å), being of the same ionic radii as Ca2+, was also more difficult to exchange because of lower electronegativity than Pb2+. Furthermore, the hydrated Cu2+ and Cd2+ ions need to be dissociated before exchanging with Ca2+. Thus, there is a possible additional rate-determining step that slows the rate of ion exchange. Further observation by Bailliez et al. noted that the Pb2+ cations have largely occupied the Ca(II) position to form Pb10-xCax(PO4)6(OH)2 solid solution[20]. In Figure 2C, the gradient of the slope is close to unity, which indicates an equimolar exchange of Pb2+ and Ca2+ cations. The Pb2+ cations are first adsorbed onto the HAP surface, especially the POH group, via surface complexation, which occurs very quickly. After complexation, Pb concentration saturates and slows down. Adsorption continues due to the high surface area of HAP powder but at a slower rate as the surface adsorption approaches equilibrium. The HAP surface may be heterogeneous, meaning some areas have lower affinities for Pb2+ cations[20]. The Langmuir-Freundlich model fits better than the Langmuir or Freundlich models, suggesting a heterogeneous HAP surface with varying affinities for cationic adsorption. This indicates two things: first, cations adsorb onto the HAP surface as a monolayer through an ion-exchange mechanism. Second, multiple active sites participate in cation adsorption and integration into the HAP lattice via cation substitution, explaining the Langmuir-Freundlich fitting results[38].
Various cations with differing ionic radii can alter unit cell dimensions when replacing calcium. For example, substituting Ca2+ with Mg2+ decreases the lattice constant c by 0.33% and increases constant a by 0.1%, leading to significant lattice disturbances and possibly more defects in HAP surfaces[39]. Introducing metal cations smaller than Ca shrinks the lattice unit cell and increases the X-ray diffraction (XRD) Bragg angle, resulting in smaller HAP size and greater irregularity shape. Cu2+ and Zn2+ preferentially substitute at the Ca(II) to form a 4 to 5-fold coordination with the HAP support[40]. Solid-state nuclear magnetic resonance (NMR) can be used to determine whether the cation substitutes at the Ca(I), or Ca(II) site. Pizzala et al. used solid-state NMR and observed chemical shift in 43Ca NMR, as compared to pure HAP phase, for different concentrations of substituting cations of Pb and V, as shown in Figure 2D[21]. Another technique to confirm cationic substitution is Raman spectroscopy. The characteristic peak belonging to PO42- can be shifted to a lower Raman shift if the Ca cation is substituted with another metal cation. For divalent cations with larger ionic radii, such as Pb2+, the substitution increases both lattice constant a and c by around 0.05% and 0.04%, respectively[35]. These changes in the crystal lattice parameters often cause changes in crystallinity, thermal stability, and morphology and may influence product selectivity in DRM and thermal catalytic CO2 hydrogenation reactions[41].
Ionic radii depend on valence electron density and the effective nuclear charge of cations. For the same nuclear charge, higher valence charge leads to smaller ionic radii due to stronger electrostatic attraction. Conversely, decreased valence charge results in larger ionic radii. Structural defects may form to accommodate other metal ions and maintain charge neutrality. HAP uniquely preserves the crystallographic symmetry of P63/m even when Ca/P drops below 1.67. Density functional theory (DFT) computational studies provide insights into HAP structure, indicating how different cations substitute in Ca(I) or Ca(II) sites[42]. Ellis et al. combined DFT with experimental methods such as X-ray absorption near edge structure (XANES) and extended X-ray absorption fine structure (EXAFS) to identify the preferred Ca site for Pb substitution, demonstrating that Pb2+ favors the Ca(II) sites[43].
For trivalent cations such as Ce3+, Fe3+, and Al3+, which have ionic radii slightly larger than Ca2+, the extent of cation substitution into the HAP lattice is again dependent on the electronegativity of the cations. The coprecipitation method was utilized to synthesize the cation-substituted materials. The authors observed morphological changes by adding different trivalent components[44]. The ones added with Fe3+ had a longer nanorod morphology than those of Al3+ and La3+. Furthermore, the addition of such trivalent elements into the HAP lattice brought about reduced crystallinity, as observed from the XRD data. This indicates that ion-exchanged HAP, be it through coprecipitation or immersion, becomes either more amorphous or has reduced particle size. Although La3+ has similar ionic radii to Ca2+, the order of ionic substitution goes in the following order: Al3+ = Fe3+ > La3+. The charge density goes in the following order: Al3+ > Fe3+ > La3+. Hence, the extent of ion exchange with such trivalent cations appears to be dependent on the component charge density. Cerium can also be substituted into the HAP lattice to create highly amorphous HAP phases[45]. Furthermore, as the ionic radius of Ce is larger than Ca2+, the lattice constants a and c are observed to increase with a corresponding increase in Ce concentration[46]. A possible reason behind the amorphousness of the Ce-HAP could be the presence of mixed Ce3+ and Ce4+ cations in the HAP lattice, which creates a higher concentration of oxygen vacancies on the HAP surface. The presence of oxygen vacancies that would be beneficial in several catalytic reactions will be discussed in later sections.
Anionic substitutability
The HAP structure contains two anionic sites located at the OH- group and the PO42- group. The most common anionic substitution is the carbonate anionic substitution. In such anionic substitution, the surface hydroxyl (OH-) is substituted with the carbonate group (CO32-), also known as A-site substitution. The A-site substitution results in the HAP chemical formula in the following manner: Ca10(PO4)6(OH)2-2k(CO3)k. Another form of anionic substitution is the substitution of the phosphate group. The phosphate group
Anionic substitution at either of these two sites can tune surface acidity. CO32- substitution at the A site can create two types of acid sites: strong acid sites from the ring of Ca2+ cations and very strong acid sites ascribed to proton contribution from the HPO42- group[48]. CO32- substitution can create lattice mismatch as CO32- is larger than OH- but smaller than PO42-[49]. As the surface acidity is altered by the presence of anionic substitution of the apatite cell structure, the catalytic activity and product selectivity can be altered. If F- is substituted into Pb-HAP, the catalytic activity for oxidative coupling of methane can be favored towards ethylene as more electrons are donated to stabilize methyl intermediate[41]. Further discussion on the effect of anionic substitution on catalytic activity will be provided in the later sections.
SUSTAINABLE CHEMICAL AND ENERGY PRODUCTION
Sustainable chemical production
Oxidative CH4 coupling to higher order hydrocarbon
Higher order hydrocarbon (OCM) has been a “holy grail” for economically viable methane conversion into ethylene and ethane for decades since Lunsford’s seminal paper[50]. Traditional OCM catalyst includes promoted silica-based systems and rare-earth oxides to give ca. 20%-30% C2 yield and around ~80 C2 selectivity[51,52]. Importantly, basicity is one key strategy to stabilize methoxy radicals with optimal binding strength for C2+ hydrocarbon formation[52,53]. In recent years, HAP-based catalysts, especially Pb-substituted HAP, with appropriate synthesis procedures and treatment, have been demonstrated to be active for OCM to produce ethane and ethylene [Table 1], to achieve around 22% C2 yield (35% CH4 conversion and 62% C2 selectivity) at 775 °C[54]. This performance is about five-fold higher selectivity than the undoped HAP, highlighting the promotional effect of Pb. Pb2+ sites doped in HAP were found to moderate active oxygen species [i.e., higher surface basicity confirmed by CO2-temperature-programmed desorption (TPD) study] and aid in stabilizing key intermediate methoxy radicals, preventing their premature oxidation to COx. In addition, Pb improves structural stability; for example, in Ca-Pb HAP, the intact apatite structure is preserved during reaction, preventing deactivation and maintaining active O- sites[55]. Other dopants have also been considered, including Sr, Ba, and Ca, but studies revealed that mono-dopants yield poor OCM activity, with the exception of Pb-substituted HAP[56,57]. The strategy is therefore shifting toward co-doping, where a secondary dopant (e.g., Pb, Cl-, or Zr) is introduced to balance surface basicity, enhance methyl radical coupling, and stabilize the apatite framework under high-temperature oxidative conditions.
Catalytic performance of HAP-based catalysts for oxidative coupling of methane (OCM)
Catalyst | WHSV (mL/gcat.h) | T (°C) | Pressure (bar) | Feed composition | CH4 conversion (%) | C2 and (CO2) selectivity (%) | Stability (mins) | Reference |
Pb2Ca8(PO4)6(OH)0.5Cl1.5 | 6,000 | 740 | 1 | P(CH4) = 21.6 kPa P(O2) = 10.8 kPa | 35 | 62(n.r.) | 3,600 | [54] |
Remarks: Chloro-HAPs were prepared with Pb substituted in HAP framework to optimize basicity, a key characteristic feature for determining OCM activity | ||||||||
Ca9.5Pb0.5 | 6,186 | 800 | 1 | P(CH4) = 21.8 kPa P(O2) = 10.9 kPa | 40 | 55(n.r.) | 2,400 | [55] |
Remarks: Hydroxyapatites with optimal Pb substitution (i.e., Ca9.5Pb0.5; Ca(10-x)Pbx(PO4)6(OH)2, x = 0.5) exhibited the best performance due to improved thermal structural integrity | ||||||||
SrC1ZrP(13/4) | 6,000 | 750 | 1 | P(CH4) = 16.5 kPa P(O2) = 8.3 kPa | 25 | 52(n.r.) | 3,000 | [56] |
Remarks: Promoted strontium chlorapatite [Sr10C12(PO4)6] has relatively thermally stable SrCl to increase the concentration of alkali beneficial for OCM | ||||||||
CaHAP-PbHAP | 3,600 | 700 | 1 | P(CH4) = 28.7 kPa P(O2) = 4.1 kPa | 18 | 80(n.r.) | 1,440 | [57] |
Remarks: CaHAP-PbHAP containing 30 wt.% PbHAP exhibited the best performance, attributed to the synergistic interaction between CaHAP and PbHAP, where CaHAP actively facilitates hydrogen abstraction from methane, while PbHAP stabilizes the resulting methyl radicals | ||||||||
Pb-HAP-CO3 | 9,200 | 700 | 1 | P(CH4) = 27.1 kPa P(O2) = 11.0 kPa | ~25 | 35(65) | 600 | [58] |
Remarks: The substitution of Pb into the HAP crystal structure stabilizes methyl radicals for C2 product formation, while the CO3 group improves the catalytic stability of the catalyst | ||||||||
Pb-HAP (0.1) | 3,000 | 700 | 1 | P(CH4) = 29.0 kPa P(O2) = 4.0 kPa | ~14.5 | ~20(20) | 180 | [59] |
Remarks: The addition of Pb was observed to improve C2 selectivity due to the stabilization of methyl radicals to form C2 products | ||||||||
Pb-HAP (c-surface) | 34,300 | 700 | 1 | P(CH4):P(O2) = 4:1, unspecified N2 dilution | ~5.0 | ~35(35) | 600 | [60] |
Remarks: The c-surface on Pb-HAP enables the more facile formation of hydroxyl vacancies which can absorb molecular O2 and make the stabilization of methyl radicals more energetically favorable. This results in higher C2 selectivity in OCM |
Oxidative propane dehydrogenation
Commercial propane dehydrogenation is an endothermic catalytic reaction that produces propylene as a valuable chemical intermediate. Thermal energy input is required to drive the catalytic reaction, incurring CO2 emissions[61]. Although the CO2 emission for commercial propane dehydrogenation is lower than steam cracking of naphtha, there is still a need to lower the CO2 emission by making the process more exothermic using oxidants such as O2 and CO2[62,63]. This makes the process more sustainable and commercially viable. However, when oxidants such as O2 are used, the propylene selectivity becomes greatly reduced and the catalyst becomes easily oxidized and sintered; thereby, deactivating the catalyst over time[64]. Hence, it is crucial to explore alternative materials for catalyst development for this catalytic reaction.
For oxidative propane dehydrogenation catalyst, V/SiO2 exhibited a reasonably good performance of 5% propane conversion and 66% propylene selectivity[65]. HAP-based catalysts displayed around 5% propane conversion and 26.4% propylene selectivity as seen in Table 2. HAP facilitates C-H bond activation through its inherent basic sites, particularly surface OH- groups and phosphate units, which promote the initial hydrogen abstraction from propane. Additionally, the incorporation of metal ions (e.g., Co2+ and Cr3+) into the HAP lattice modifies its redox and acid-base properties. For example, Co-doped HAP facilitates hydrogen abstraction via surface OH- group[66], whereas Cr-substituted HAP enhances release of lattice oxygen for hydrogen abstraction[67]. The propylene selectivity is still lower than the V/SiO2 benchmark catalyst. Hence, more research effort is required to enhance propylene selectivity through a more rational understanding of the catalyst surface.
Catalytic performance of HAP-based catalysts for oxidative propane dehydrogenation
Catalyst | WHSV (mL/gcat.h) | T (°C) | Pressure (bar) | Feed composition | C3H8 conversion (%) | C3H6 selectivity (%) | Stability (mins) | Reference |
Co55SrHAP | 3,600 | 450 | 1 | P(C3H8) = 14.5 kPa P(O2) = 4.1 kPa | 22.1 | 48.6 | 360 | [66] |
Remarks: The integration of Co2+ into the Sr-HAP lattice encourages hydrogen abstraction due to easier removal of H from the surface OH- group | ||||||||
Cr(3.7)/CaHAP | 90,000 | 550 | 1 | P(C3H8) = 6.08 kPa P(O2) = 3.04 kPa | ~20.0 | ~34.0 | 300 | [67] |
Remarks: The Cr species on the HAP surface become more oxidized after calcination in the presence of air into Cr6+. Cr6+ was eventually reduced to Cr3+ during the reaction. The presence of Cr3+ favors the release of lattice oxygen which is important for hydrogen abstraction for propylene formation | ||||||||
V-CaHAP (V/P = 15) re-oxidized | 3,600 | 450 | 1 | P(C3H8) = 14.5kPa P(O2) = 4.1 kPa | 15.2 | 55.2 | 360 | [68] |
Remarks: The redox cycling between V4+ and V5+ and lattice oxygen (abstraction and incorporation) are important in propane conversion and propylene formation | ||||||||
Ca10(PO4)0.78(VO4)5.22(OH)1.32O0.34 | 7,500 | 450 | 1 | P(C3H8) = 7.5 kPa P(O2) = 2.13 kPa | ~5.0 | ~27.0 | 90 | [69] |
Remarks: The formation of vanadium oxy-HAP created basic sites that are essential for activating the C-H bonds in propane molecules |
Biodiesel production
Diesel is derived from petroleum distillation and is widely used in vehicle transportation. Biodiesel production from renewable sources is strongly considered a more sustainable pathway than oil-derived diesel. Biodiesel comprises fatty acid alkyl esters (FAME) which can be synthesized from vegetable oil or animal fats via a transesterification reaction with methanol as the reactant[70]. Waste cooking oil can be used as an alternative feedstock which lowers biodiesel production costs[71]. Basic homogeneous catalysts such as NaOH and KOH are effective for biodiesel production but pose challenges such as corrosiveness, difficult separation, and wastewater generation[72], while solid CaO offers recyclability and high yields, but suffers from deactivation by CO2/H2O and leaching due to reaction with glycerol[73,74]. HAP-based catalysts demonstrate several innovations that address the key limitations of conventional CaO-based systems in biodiesel production. Notably, HAP provides excellent surface basicity and structural stability, enabling a high biodiesel yield of 95% for eight cycles with negligible leaching as seen in Table 3; hence, HAP-based catalysts deserve closer research attention[75].
Catalytic performance of HAP-based catalysts for biodiesel production
Catalyst | T (°C) | Feed condition | Conversion (%) | No of cycles | Reference |
30K/HAP-600 | 65 | 30 g of palm oil and an appropriate amount of methanol | 96.4 | 8 | [75] |
Remarks: The synergistic effect of K2CO3 addition and calcination at 600 °C maximized surface basicity; thus, maximized FAME yield for 8 reaction cycles at ca. 90%. Slight deactivation occurred due to K+ leaching | |||||
Ni-Ca-HAP | 70 | The waste fish scale has a high acid value | 59.9 | 2 | [76] |
Remarks: The Ni-Ca-HAP solid acid catalyst can be applied to convert high acid value feedstock into biodiesel (FAME) compared with the calcined waste fish scale (WFS) "base catalyst". The Design of Experiment methodology was applied and revealed that methanol flowrate and nickel nitrate loading are the most important parameters in maximizing FAME yield | |||||
11 wt.% dosage of 30%CaO-CeO2/HAP-650 | 65 | 30 g of palm oil with a 9:1 methanol to palm oil molar ratio | 91.84 | 8 | [77] |
Remarks: The synergistic addition of CaO-CeO2 onto bone-derived HAP enhanced FAME yield in 8 reaction cycles due to higher basicity. The leaching concentration dropped over more cycles due to lattice distortion by Ca2+ and Ce4+ substitution. The specification of the produced biodiesel is within the ASTM standard for biodiesel | |||||
5%CaO-HAP | 60 | 100 g of vegetable oil | 95.18 | 4 | [78] |
Remarks: The bone-derived HAP support possessed a higher BET area than pure CaO, even though CaO is postulated as the active site. The extent of CaO penetration into the HAP pores or CaO dispersion on the HAP surface may be evaluated using visual characterization techniques such as TEM | |||||
50-NaHAP-800 | 100 | Methanol to oil molar ratio of 6:1 | 99 | 5 | [79] |
Remarks: Sodium nitrate loading and calcination temperature were found to be important in maximizing FAME yield by enhancing surface basicity. The produced biodiesel meets EN14214 standards | |||||
HAP-γ-Fe2O3 | 65 (methanol reflux temperature) | Methanol/soybean oil molar ratio of 25:1 | 94 | 5 | [80] |
Remarks: HAP was encapsulated around γ-Fe2O3 via the coprecipitation method to form a core-shell nanoarchitecture before immobilizing with biguanide molecules. The soybean conversion to methyl ester is close to 100% for 5 cycles with little deactivation. The catalyst can be recycled due to the strong magnetic properties of the γ-Fe2O3 core | |||||
Marble slurry-based HAP | 65 | Methanol/soybean oil molar ratio 9:1 | 94 | 5 | [81] |
Remarks: HAP is found to be a better catalyst than calcined marble slurry (CMS) for biodiesel production in terms of biodiesel yield and fuel properties because of higher basicity | |||||
HAP | 65 | The canola oil, rapeseed oil, and waste cooking oil to methanol molar ratio was taken as 1:12 | 89.4%, 96.7% and 91.7% were achieved for canola oil, rapeseed oil, and waste cooking oil, respectively | - | [82] |
Remarks: The synthesized catalyst showed high total basicity after calcination |
Hydrogen production
Dry reforming of methane (DRM)
Catalytic transformation of greenhouse gases such as CO2 and CH4 to produce syngas is highly desirable as syngas can be used as feedstock for the Fischer-Tropsch process, methanol synthesis, and one-step dimethyl ether (DME) synthesis process[83]. Thus, it can be considered a sustainable path for syngas production. Syngas can also be used as a source of hydrogen or directly to produce electricity[84,85]. DRM, which uses CO2 as a feedstock, is a desirable alternative to current syngas generation processes from an environmental and sustainability perspective[86]. A major technical challenge in DRM is rapid catalyst deactivation induced by severe coke formation during the reaction, which hinders the commercial development of the DRM process[87]. The last decade has thus seen vigorous research in catalyst development for this process to increase activity, stability, and resistance to coking[88].
Conventionally, Ni/MgAl2O3 catalysts were commonly used as a benchmark, which exhibited 90% CH4 conversion and approx. 92% CO2 conversion at
Catalytic performance of HAP-based catalysts for dry reforming of methane
Catalyst | WHSV (mL/gcat.h) | T (°C) | Pressure (bar) | Feed composition | CH4 conversion/CO2 conversion (%) | H2/CO ratio | Stability (h) | Carbon formation (%wt.) | Reference |
2Ni1/HAP-Ce | 60,000 | 750 | 1 | 20 mol%CH4 20 mol%CO2 60 mol%He | ~95/80 | ~0.8 | 100 | ~0 | [89] |
Remarks: Ce-doping induces strong metal-support interaction which anchors the Ni cluster and single atoms and activates the C-H bond in methane. This results in high activity and stability for 100 h with near negligible carbon formation | |||||||||
Ni0.5/HAP-OP | 30,000 | 800 | 1 | 10 mol%CH4 10 mol%CO2 80 mol%N2 | 93.3/98.8 | 1.0 | 200 | ~0 | [90] |
Remarks: The coprecipitated Ni/HAP catalyst displayed very high reactant (CO2 and CH4) conversion due to the quick conversion of CH4 into carbon and H2. The carbon gasification into CO is equally quick, resulting in negligible carbon accumulation on the catalyst surface. The surface basicity of the catalyst requires more careful investigation | |||||||||
Ca-HAP2 | 15,900 | 700 | 1.6 | 20 mol%CH4 20 mol%CO2 60 mol%N2 | 78/84 | n.r. | 90 | n.r. | [91] |
Remarks: The Ni-HAP catalysts with different surface areas displayed comparable catalytic activity and stability with existing literature of that time due to small Ni nanoparticle size, surface basicity, and large surface area (66 m2/gcat). But no results on carbon formation and H2/CO ratio for the two catalysts were shown | |||||||||
Co-HAP-N | 31,764 | 700 | 1.6 | 20 mol%CH4 20 mol%CO2 60 mol%lN2 | 75/55 | ~0.75 | 50 | ~10 | [92] |
Remarks: The non-stoichiometric Co-HAP catalyst performed better than the stoichiometric Co-HAP. The carbon formation for the NiCo-HAP catalyst was lower than the Co-HAP or Ni-HAP catalyst | |||||||||
Ni(4)/CaHAP | 38,400 | 600 | 1 | 3 mol%CH4 3 mol%CO2 94 mol%lN2 | 78/78 | ~0.70 | 4 | ~50 | [93] |
Remarks: Three types of Ni species were observed on the surface using H2-Temperature Programmed Reduction (TPR). A Higher amount of Ni2+ was substituted into the HAP lattice at lower Ni loading. With the increasing amount of Ni loading, more NiO is formed and therefore less Ni2+ is available for cation substitution into the HAP lattice. Carbon formation correspondingly increases with Ni loading | |||||||||
Ni-Co/HAPSIWI | 8,823 | 750 | 1.6 | 20 mol%CH4 20 mol%CO2 60 mol%N2 | 73/79 | ~0.90 | 160 | < 13 | [94] |
Remarks: Bimetallic Ni-Co species were present and cation-exchanged onto the HAP support for good and stable catalytic performance. A limited amount of carbon was observed on the catalyst surface using TEM characterization but was not quantified using the TGA technique | |||||||||
Pt-HAP | 8,400 | 700 | 1 | 20 mol%CH4 20 mol%CO2 60 mol%N2 | 40/30 | ~0.81 | 50 | n.r. | [95] |
Remarks: Ru and Pt-based HAP catalysts were prepared using different deposition methods such as incipient wetness impregnation method, cation exchange, and excess liquid impregnation. The one prepared by incipient wetness impregnation displayed reasonable catalytic performance over time than other methods. A water trap was used to quantify water formation over time | |||||||||
0.5Ni1/HAP-SAC-PVP | 60,000 | 750 | 1 | 20 mol%CH4 20 mol%CO2 60 mol%He | ~75/85 | ~0.9 | 16 | ~0 | [96] |
Remarks: The Ni-HAP assisted by PVP addition exhibited good catalytic performance over time due to the greater prevalence of Ni single atoms. The authors did not further expound on the role of PVP in enhancing the catalytic performance and stability of the Ni single atom on the HAP catalyst surface | |||||||||
Ni/HAP-80 | 24,000 | 750 | 1 | 50 mol%CH4 50 mol%CO2 | 44/55 | ~0.77 | ~12 | 0.99 | [97] |
Remarks: The ratio of mesopores and macropores of HAP can be controlled when the heating temperature varies in the presence of cetyltrimethylammonium bromide (CTAB). This ratio can influence the Ni particle size which can affect catalytic performance and stability | |||||||||
Ni/Ca-HA1_S | 15,882 | 700 | 1 | 50 mol%CH4 50 mol%CO2 | 50/60 | 0.8-0.9 | 30 | < 10 | [98] |
Remarks: The regeneration property of the Ni-HAP catalysts in different oxidizing atmospheres (O2 and CO2) was investigated. There is a very small decrease in the CH4 conversion of the Ni-HAP catalyst over three cycles. The reason for the decrease in catalytic activity could be Ni nanoparticle sintering during the reaction |
Partial oxidation of CH4 (POM)
The current technology for syngas production, methane steam reforming (ΔH°298 = 206.0 kJ/mol), requires a tremendous amount of energy to break the C-H bond in methane and the O-H bond in steam. As a result, the CO2 emission per hydrogen produced (i.e., 11.5 tons CO2 emitted per ton H2) is the highest among the available technologies, such as natural gas-fired methane pyrolysis and electric-arc furnace methane pyrolysis for hydrogen production and creates a strong motivation to move towards a more sustainable hydrogen production[100].
POM is an attractive alternative to the commercial steam reforming of methane reaction as it is an exothermic reaction, and a high H2/CO ratio (~2) can be attained. Higher H2/CO is advantageous for Fischer-Tropsch, solid oxide fuel cells, and methanol synthesis from CO hydrogenation[101]. The technical issues with POM are that full methane combustion can occur as a side reaction, which can reduce the
Catalytic performance of HAP-based catalysts for partial oxidation of methane (POM)
Catalyst | WHSV (mlCH4/gcat.h) | T (°C) | Pressure (bar) | Feed composition | CH4 conversion (%) | CO selectivity (%) | Stability (mins) | Reference |
Ca90NiP(25) | 4,800 | 800 | 1 | P(CH4) = 16.2 kPa P(O2) = 8.1 kPa | ~95.0 | ~95.0 | n.r. | [102] |
Remarks: The hysteresis in CH4 conversion can be attributed to CH4 activation by Ni0 as the temperature increases. But more of the Ni0 species become oxidized into NiO by O2 when temperature decreases. The NiO can still be partially reduced by reactant gas into a mixture of Ni0 and NiO which makes the catalyst more difficult to reduce but more catalytically stable. The spent Ni-Ca-HAP exposed to ambient air could be reactivated under standard reaction conditions | ||||||||
HAP(1.61) | 3,000 | 600 | 1 | P(CH4) = 29.0 kPa P(O2) = 4.0 kPa | ~6.0 | n.r. | 180 | [103] |
Remarks: The formation of PO43- ions on the non-stoichiometric HAP surface appears to be favorable to the formation of CO due to enhanced CH4 activation. As the Ca/P ratio decreases, higher formaldehyde formation was observed | ||||||||
SrHAP | 3,600 | 600 | 1 | P(CH4) = 28.7 KPa P(O2) = 12.3 kPa | ~10 | ~90.0 | 5,050 | [104] |
Remarks: The SrHAP attained high selectivity toward CO formation in the initial 6 h but the CH4 decreased beyond that. The reason for the decrease in catalytic activity is due to the formation of | ||||||||
Pb12HAP | 3,600 | 600 | 1 | P(CH4) = 28.7 kPa P(O2) = 4.3 kPa | ~11.0 | ~30.0 | 180 | [105] |
Remarks: The partial Pb substitution into the HAP lattice was observed to be beneficial for CH4 activation. The addition of tetrachloromethane (TCM) could chlorinate the surface and cause catalyst deactivation. However, the CH3Cl formation was also observed | ||||||||
BaHAP | 3,600 | 700 | 1 | P(CH4) = 28.7 kPa P(O2) = 4.1 kPa | ~8.0 | ~10.0 | 6 | [106] |
Remarks: Ba addition was found to have beneficial effects on CH4 conversion for POM in the absence of TCM. The introduction of TCM can be beneficial for C2 selectivity but detrimental to catalyst stability. Methyl chloride formation was also observed | ||||||||
HAP | 3,600 | 700 | 1 | P(CH4) = 28.7 kPa P(O2) = 4.1 kPa | ~4.0 | ~10.0 | 6 | [107] |
Remarks: CH4 conversion and C2 selectivity do not significantly change with Ca/P ratio and time. The introduction of N2O and on-stream TCM can be detrimental to CH4 conversion and C2 selectivity | ||||||||
Ni-Sr-HAP | 6,000 | 750 | 1 | P(CH4) = 16.1 kPa P(O2) = 8.2 kPa | ~90.0 | ~90.0 | n.r. | [108] |
Remarks: Although the catalyst subjected to the shortest calcination time of 2 h took a lower temperature to activate CH4 molecules, the CH4 conversion for the other two catalysts at different calcination times is very similar beyond ~780oC. Longer calcination time creates more Sr3-xNix (PO4) 2 less reducible phase | ||||||||
Ca95NiP(25) | 4,800 | 750 | 1 | P(CH4) = 16.1 kPa P(O2) = 8.2 kPa | ~80.0 | ~88.0 | 84 | [109] |
Remarks: The CaNiP catalysts were prepared by the coprecipitation method and the Ca/PO4 ratio was found to be optimal for CH4 conversion at 10/6 due to the presence of surface basic sites and reducible Ni sites for CH4 activation | ||||||||
Ce0.1Ni2.5Ca10 | 32,000 | 600 | 1 | P(CH4) = 16.1 kPa P(O2) = 8.1 kPa | ~65.0 | ~60.0 | 20 | [110] |
Remarks: Ce, added as a promoter after the coprecipitation was completed, could enhance CH4 conversion and improve catalyst stability as Ce can provide oxygen vacancies to promote carbon gasification and so reduce carbon formation | ||||||||
Ni-Ca-HAP | n.r. | 750 | 1 | P(CH4) = 16.2 kPa P(O2) = 8.1 kPa | ~90.0 | ~85.0 | n.r. | [111] |
Remarks: The Ni-Ca-HAP catalyst was subjected to N2O pulse chemisorption. The adsorbed CH4 is converted into CO2 due to lattice oxygen on the HAP surface. When lattice oxygen depletes, CO formation and, subsequently, carbon formation become more favored. Partially oxidized Ni species are required for high CH4 conversion and stable performance. The catalyst was deposited onto monolith walls, and it was found that at higher temperatures, the temperature change decreases due to higher heat loss at the reactor walls | ||||||||
Rh(1)-HAP | 19,200 | 700 | 1 | P(CH4) = 10.1 kPa P(O2) = 5.1 kPa | ~76.0 | ~85.0 | n.r. | [112] |
Remarks: Three Rh species were observed on the HAP surface from characterization techniques: surface Rh particles, RhOx with good interaction with the support, and Rh incorporated into the HAP lattice structure. Although the dispersion of Rh (1)-HAP appears relatively low, the catalyst stability remains well-preserved over 30 h due to better metal-support interaction which makes it difficult to sinter and oxidize. Rh (1)-HAP catalyst appears comparable to industrial Rh-Al2O3 catalyst in CH4 conversion and H2 yield |
CO oxidation
CO oxidation is an integral reaction to eliminate CO in hydrogen fuel applications and emission control in automobile applications to meet government emission standards[113]. Hence, it is an important reaction that can speed up the use of hydrogen fuel cells and make transportation more sustainable. Typically, noble metals such as Au have shown to be effective in creating highly active sites for CO oxidation due to the formation of cationic Au and metallic Au, and this may enhance O2 dissociation at the metal-support interface[114-116]. The HAP-based catalyst is seven times more active than the benchmark catalyst (Au/Fe2O3) (i.e., 0.071 molCO/gAu·h vs. 0.011 molCO/gAu·h) [Table 6][117]. The excellent catalytic performance indicates that HAP-based catalysts deserve more careful investigation to develop an active and stable catalyst for this reaction
Catalytic performance of HAP-based catalysts for CO oxidation
Catalyst | WHSV (mL/gcat.h) | T50% (°C) | Pressure (bar) | Feed composition | Stability (h) | Reference |
Au/FH-400 | 20,000 | ~-25 | 1 | 1.0 mol%CO 1.0 mol%O2 98 mol%He | 300 | [117] |
Remarks: The addition of FeOx, together with increasing calcination temperature, creates strong metal-support interaction for a very stable catalytic performance of 300 h | ||||||
Au/HAP-O2 | 100,000 | 14 | 1 | 1 mol%CO 20 mol%O2 79 mol%He | 24 | [118] |
Remarks: The calcination atmosphere of the Au/HAP catalyst influences the catalytic activity and stability. Although calcination in the He atmosphere produces very fine Au nanoparticles, the rate of deactivation is also the sharpest. However, the one calcined in the O2 atmosphere exhibits high catalytic activity over 24 h. Strong metal-support interaction between Au and HAP support was proposed to be the main reason | ||||||
Au/HAP | 31,500 | 35 | 1 | 3.4 mol%CO 21 mol%O2 75.6 mol%He | n.r. | [119] |
Remarks: The HAP pre-treatment temperature influences CO conversion with nearly 100% conversion at room temperature for the one calcined at 300 °C. The interaction between CO and the HAP support creates structural vacancies that enhance catalytic activity | ||||||
Au/HAP(400) | 20,000 | ~0 | 1 | 1.0 mol%CO 1.0 mol%O2 98 mol%He | n.r. | [120] |
Remarks: The strong metal-support interaction (SMSI) effect observed in Au/HAP is reversible, except that it occurs in oxidative conditions. This is the opposite of classical SMSI. SMSI effect helps to make the Au nanoparticle more sintering resistant | ||||||
Au-FH(400)-17.5% | 20,000 | ~-30 | 1 | 1.0 mol%CO 1.0 mol%O2 98 mol%He | n.r. | [121] |
Remarks: The presence of FeOx helps to reduce carbonate, an inactive reaction intermediate species, accumulation which helps to enhance catalytic activity | ||||||
Au-Cu-HAP | 10,890 | ~30 | 1 | 10 mol%CO 90 mol%air | 14 | [122] |
Remarks: The addition of Au created more O2 adsorption sites and increased Cu dispersion on the HAP support. The higher Cu dispersion appears to tune the Cu valence between +1 and +2. The latter effect enables the catalyst to be active and stable as compared to a monometallic Cu-HAP catalyst | ||||||
Cu(4)-HAP | 80,000 | ~175 | 1 | 1.0 mol%CO 1.0 mol%O2 98 mol%He | n.r. | [123] |
Remarks: The Cu-HAP catalysts with lower loading (0.8% and 3.6%) exhibited better catalytic performance. The Cu lattice micro-strain correspondingly increases with Cu loading and appears to contribute catalytic activity for CO oxidation |
Water gas shift reaction
Currently, the WGS reaction is employed to increase the H2/CO ratio of syngas produced from the steam methane reforming (SMR) process. Thereby enhancing the sustainability of the SMR process. The commercial WGS catalysts used are the Fe-Cr mixed oxide catalyst and the Cu-Zn mixed catalyst used at a high-temperature range (350-450 °C) and a lower temperature range (190-250 °C), respectively[124]. As the WGS reaction is restricted by thermodynamic equilibrium, the commercial catalytic process typically performs the reaction at high temperatures, followed by a lower temperature range to increase CO conversion and H2 yield, thus approaching thermodynamic limits. Although the Fe-Cr mixed oxide catalyst displayed moderate catalytic performance suitable for industrial purposes, it is important to note that there is strong regulatory pressure to replace the Fe-Cr mixed oxide catalyst as Cr6+ is highly carcinogenic and can cause immense harm to human life and the natural environment[125]. These concerns have motivated intense research to explore alternative materials that exhibit high catalytic performance and yet are less harmful to human life and the natural environment. A work by Miao et al. demonstrated that HAP-based catalysts can display better catalytic performance than CeO2-based catalysts in WGS, as a consequence of the unique HAP structure where H2O could be polarized by Lewis acidic Ca2+ ions and H bonding to basic O atoms of PO43- units[126]. Thus, the HAP-based catalyst deserves detailed investigation to improve WGS catalytic performance.
CO2 utilization
CO2 hydrogenation to CH4
As the mitigation of CO2 emissions becomes a pressing issue globally, CO2 utilization technologies via thermal and photothermal means have gathered momentum in recent years[127]. CO2 hydrogenation to methane (i.e., CO2 methanation) has gained increasing attention as an alternative method to produce sustainable methane while reducing CO2 emissions at the same time[128-130]. HAP has gained traction as an alternative support in CO2 methanation[131]. As photothermal CO2 reduction is still in a nascent stage of development, significant effort is required to develop an active, stable, and selective catalyst before scaling-up efforts can be undertaken[132]. The Cu/HAP catalyst demonstrates catalytic performance on par with other metal oxide-based catalysts, indicating that HAP-based catalysts warrant further investigation for photothermal CO2 reduction[133]. The catalytic performance of the HAP-based catalysts for CO2 methanation can be seen in Table 7.
Catalytic performance of HAP-based catalysts for CO2 hydrogenation to CH4
Catalyst | WHSV (mL/gcat.h) | T (°C) | Pressure (bar) | Feed composition | CO2 conversion (%) | CH4 selectivity (%) | Stability (h) | Reference |
NiHAP (0.5) | 30,000 | 350 | 1 | 8 mol%CO2 32 mol%H2 60 mol%He | 75-80 | 97.0 | 70 | [134] |
Remarks: The addition of oleic acid was found to influence the chemical state of Ni which can affect the amount of surface formate on the catalyst surface. A higher amount of Ni0 is important for methane formation as CO desorbs easily from partially reduced Ni sites | ||||||||
Ni/La-(6.6) | 30,000 | 350 | 1 | 16 mol%CO2 64 mol%H2 20 mol%He | 70 | 98.0 | 100 | [135] |
Remarks: La acts as a dual-function material: the first purpose is to chemisorb CO2 while the second purpose is to create highly dispersed Ni nanoparticles supported on HAP. The presence of such nanoparticles provides active sites for active and stable catalytic CO2 methanation | ||||||||
10Ni6CeHA | 12,000 | 325 | 1 | 8 mol%CO2 32 mol%H2 60 mol%He | 92.5 | 100 | n.r. | [136] |
Remarks: Ce was added onto the Ni/HAP catalyst surface and increased the catalytic activity and methane selectivity. The improved catalytic performance can be attributed to increased oxygen vacancy and higher dispersion of the Ni nanoparticles |
SURFACE ENGINEERING OF HYDROXYAPATITE CATALYSTS
Active metals deposition
The active center affects the reaction rate, selectivity, and stability of a catalytic reaction. Active centers can be introduced to HAP support through several methods, with incipient wetness impregnation as the commonly used strategy. Alternatively, electrostatic adsorption using ammonia can create dispersed active sites on the HAP surface. Additionally, coprecipitation and ion-exchange methods can enhance surface basicity and create dispersed active sites on the HAP catalyst surface[16].
The wetness impregnation method not only deposits the active center onto the catalyst surface but also may perform partial cation exchange as the cations in the impregnation solution may swap with the Ca2+ cation during the impregnation process[137]. This creates two different sites with different sizes: one with nanoparticles and the second one with cluster or single atom size. The active center of two different sizes may affect product selectivity. Domínguez et al. used HAP as catalyst support, impregnated with Au, for CO oxidation. This catalyst has demonstrated near complete CO conversion at room temperature due to the presence of structural defects by eliminating surface carbonates which alter the Au oxidation state[119]. Wang et al. reported that the Au/HAP catalyst calcined under reaction conditions is more active than in air[138]. Higher Au loading resulted in lower CO activity. Further characterization using X-ray photoelectron spectroscopy (XPS) revealed that higher HAP crystallinity resulted in a lower Au3+/Au0 ratio due to lower Ca/P. This indicates that the surface becomes more acidic, thereby withdrawing more electrons from the Au species and creating Au3+. Further investigation by Huang et al. has shown that the calcination temperature and calcination environment can influence CO oxidation activity and selectivity at room temperature due to the formation of very small Au nanoparticles[118]. Although very small Au nanoparticles, thought to contribute to high initial activity, have been observed in TEM images, SMSI is believed to underlie this influence. Tang et al., through an in-situ Diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) study of catalyst samples prepared at different temperatures, have indeed shown that the SMSI effect has a detrimental effect on CO oxidation activity at high calcination temperature as the active sites were encapsulated by HAP support under an oxidative environment[120]. However, such SMSI effect can be tuned and manipulated through partial encapsulation to prevent metal sintering and yet allow selective and adequate catalytic activity to occur on the active sites. In one such demonstration, Fe was introduced into the HAP lattice through cation substitution using the coprecipitation method; the Fe-HAP surface with abundant OH- groups and/or PO42- groups was speculated to anchor the Au nanoparticles, akin to dispersing Au on FeOx[139,140]. Consequently, the Au/Fe-HAP (Au/FH) catalyst is significantly more active and stable than that of the conventional Au/Fe2O3 (Au/F) catalyst, as shown in Figure 3A[117].
Figure 3. (A) Stability curves in CO oxidation of various catalysts calcined at 400 °C or dried at 60 °C. CO:O2:He = 1:1:98,
Bimetallic nanoparticles on the catalyst surface often induce two catalytic phenomena: electronic and geometric effects[141]. Guo et al. synthesized Au-Cu/HAP for CO oxidation and observed that Cu addition causes a reduction of Au nanoparticle size which resulted in greater Au dispersion. XPS characterization indicated the presence of Cu+ and Cu2+ which appeared to indicate strong interaction with Au[122]. Further characterization using O2-TPD shows that more oxygen sites are present in the Au-Cu system as compared to the monometallic Au catalytic system[122]. Other than Au, Cu has been used as an active metal for CO oxidation. Different Cu loadings had been impregnated onto stoichiometric HAP for CO oxidation and preferential CO oxidation. The lowest Cu loading at 0.8%wt Cu shows the highest catalytic activity among the other catalyst samples. Surface characterization using XPS and H2-TPR revealed the presence of Cu+ species and Cu2+ species that might be integrated into the HAP framework. Higher Cu dispersion on the lowest loading catalyst resulted in better CO chemisorption[123].
A noble metal component, such as Pd, has been used for full methane oxidation. In the presence of strong metal-support interaction - induced by increasing the calcination temperature, the Pd coordination changes from tetrahedral to square planar geometry while containing a mixture of Pd0 and Pd2+[142]. Stronger metal-support interaction can help stabilize the Pd nanoparticles in successive oxidation cycles and provide better catalytic conversion via improvement in PdO reduction[142]. Nickel has been added as a bimetallic component to enhance the catalytic performance but the monometallic nickel displays reasonable performance for full methane oxidation without carbon formation[143].
The presence of Pb appears to increase the ethylene selectivity, which transforms POM into OCM. Other than lead (Pb), barium (Ba) is also used as an active component for the POM reaction. The addition of Ba produces better methane conversion than Ca or Pb, and the resulting trend is similar to what is reported[106].
This property can be utilized to load transition metals, such as Ni, Co, etc., with strong interaction with the HAP support. Depending on the metal loading, the distribution of various types of transition metal species interacting with different support sites is obtained in HAP-based catalysts. For example, in a study by Boukha et al.[93], the Ni loading on HAP catalysts varied from 1% to 10% and it was observed that Ni was mostly present as Ni2+ ions by ion exchange with Ca2+ at lower Ni loadings (< 1 wt%). The progressive increase of Ni loading from 1% to 10% led to the formation of small NiO nanoparticles at less than 2%-4% loading and large NiO nanoparticles at higher loadings. The strong interaction of the metal with the support is highly desirable for DRM to prevent metal sintering at the high reaction temperature and subsequent formation of coke catalyzed by large metal ensembles[144,145]. However, while the ion-exchanged Ni2+ has a very strong interaction with HAP support, they exhibited low activity for DRM (as evidenced by poor conversion on the 1% Ni loading catalysts). Thus, the Ni/HAP catalyst with an intermediate Ni loading of 4% and small NiO nanoparticles showed the best performance. The poor activity of the ion-exchanged Ni species may be caused by the cationic state of Ni. Metallic Ni has traditionally been recognized as the active site for DRM, which is why a pre-reduction step before the reaction increases catalytic effectiveness activity. Ion-exchanged Ni2+ on HAP is difficult to reduce, possibly leading to lower activity than NiO nanoparticles on HAP, as also reported by other studies[91,94]. Rêgo De Vasconcelos et al. compared three loading techniques - ion exchange, incipient wet impregnation, and liquid phase impregnation for Pt and Ru on HAP support and also observed that incipient wetness impregnation resulted in better catalytic performance in DRM[95]. This broadly agrees with the higher activity of smaller metal nanoparticles on HAP support with intermediate metal-support interaction than strongly interacting ion-exchanged cationic metal species or weakly interacting larger metal nanoparticles. Similar results were also reported for Pd loaded on HAP in DRM[146].
To further increase methane conversion for POM, Ni was used as an active metal and the CH4 conversion for Ni-HAP catalyst was comparable to hydrotalcite catalyst as aforementioned[102]. Upon increasing reaction temperature, NiO nanoparticles on the HAP surface were partially externally reduced to Ni0 in the presence of reductive gases (i.e., H2, CO, CH4), along with Ni exsolution from the HAP structure [Figure 3B]. Upon cooling, the catalyst surface reoxidizes, forming a layer of NiO, and the exsolved Ni remains exsolved. When heated again, it returns to the reduced state, allowing for repeated cycles of temperature changes [Figure 3C]. The interesting catalytic performance is attributed to SMSI, where HAP support decorates Ni nanoparticles, enhancing their sintering resistance. During the reaction, larger nickel nanoparticles oxidize, while smaller ones maintain their metallic character. Fresh samples do not show nickel nanoparticles, but they appear post-reaction, suggesting that smaller nanoparticles are key for CH4 activation and H2 production[102]. In another example, Ni nanoparticles are deposited onto HAP support using oleic acid. Through the use of in-situ DRIFTS and ex-situ XANES, the amount of surface formate species is related to the chemical nature of Ni on the catalyst surface, as seen in Figure 3D[134]. A reaction mechanism for the CO2 methanation was proposed and validated using non-linear kinetic modeling. Although the use of HAP-based material can bring about improvement in CO2 conversion and CH4 selectivity, the use of more basic promoters can bring about even greater improvement in catalytic performance.
Noble metals such as Rh were also used as active metals for the POM. Three different types of Rh species were observed from H2-TPR: Rh2O3 on the HAP catalyst surface and small RhOx particles with SMSI and Rh2+ cations integrated into the HAP framework. The metal-support interaction of the active metal component determines the number of metallic sites available for POM; thus, optimum metal-support interaction is required for good methane activation and catalytic stability. Furthermore, the Rh/HAP catalyst at low Rh metal loadings (i.e., 1 wt.%) gave comparable performance to Rh/Al2O3 catalyst due to the abundance of basic hydroxyl sites and higher pore size, as shown in Figure 3E[112]. In summary, active metal deposition, the resulting metal-support interactions, and the dispersion state of metal species on HAP significantly influence catalytic performance, product selectivity, and stability across various catalytic conversion reactions.
Cationic substituted catalysts
The substitution of the Ca cations with different cations can affect catalytic activity, product selectivity, and catalyst stability for a myriad of catalytic reactions. The active sites in cation-substituted HAP are more sintering-resistant, as these active sites have better metal-support interaction as compared to the catalyst prepared by the conventional incipient impregnation method[90]. The free metal oxide nanoparticles on HAP support tend to sinter into larger particles during reduction and catalysis, leading to catalyst deactivation and side products. Substituting a transition metal into the HAP lattice improves active site dispersion, reducing side product formation[147]. The partial substitution of the Ca cations has been demonstrated for PDH. Partial cationic substitution using Co2+ has led to greater improvement in propane conversion and propylene selectivity as compared to Sr3+ and Ba2+. The partial Co substitution in the cation site altered the surface property by making OH- groups more amenable to hydrogen abstraction from propane, thus creating O- species[66].
Alternatively, chromium (Cr) can be cation-substituted into the HAP framework via the ion-exchange method and has shown interesting catalytic activity for oxidative propane dehydrogenation. The Cr species changed to Cr6+ after air calcination due to oxidation. However, the initial high catalytic activity decreases over time due to the reduction of Cr6+ species to Cr4+[67]. However, as chromium can be carcinogenic, other active metals can be explored as an alternative to chromium-based HAP catalysts for oxidative propane dehydrogenation.
CO2 reduction via the photothermal route has attracted attention for utilizing sunlight to drive catalytic reactions. Photons generate electron-hole pairs or high-energy carriers through localized surface plasmon resonance, which can also induce local heating and enhance catalytic activity through thermal effects[148]. Guo et al. have exploited the ability of HAP for partial cation substitution to create highly dispersed Cu species on the HAP surface for photothermal CO2 reduction to CO by using the coprecipitation method for catalyst synthesis[133]. The activation energy of the CO2 reduction reaction is lowered by 27.5% and CO formation during the photothermal reaction appears to be dependent on the Cu loading as observed in Figure 4A and B. The effectiveness of the photothermal approach can be attributed to the presence of photogenerated electrons and holes that could be trapped at the Lewis-acidic Cu2+ and Lewis-basic PO3-4 sites of surface frustrated Lewis pair (SFLPs) upon absorption of solar light. This consequently lowered the activation barrier of the CO2 reduction reaction[149].
Figure 4. (A) Capillary flow reactor results for the calcined 10 mol% Cu-HAP catalyst. The conversion rates and activation energies, recorded under dark and light conditions, for the selective hydrogenation of CO2 to CO are plotted. Reproduced from Ref.[133] with permission from American Chemical Society. (B) Photocatalytic reverse-WGS activity of calcined Cu-HAP samples as a function of Cu content in a batch reactor. Reaction conditions: H2/CO2 ratio = 1:1, light intensity = 40 suns, no external heating, and measurement time = 1 h. Reproduced from Ref.[133] with permission from American Chemical Society. (C) CH4 conversion for Ni/xLa-HAP catalysts. Reaction condition: CH4:CO2 = 1:1, catalyst weight = 100 mg, total flow rate = 40 mL/min, reaction temperature = 750 °C, GHSV = 24,000 mL/g·h. Adapted from Ref.[150] with permission from Elsevier. (D) Schematic of possible reaction pathways in DRM for La-substituted and pristine Ni/HAP. Adapted from Ref.[150] with permission from Elsevier.
Trivalent La3+ has been successfully incorporated into HAP up to 7 wt.% due to the close similarity in ionic radius between La3+ and Ca2+ (i.e., 101 and 100 Å, respectively)[150]. Subsequently, Ni was impregnated, and the HAP with La3+-substitution was found to exhibit better activity in DRM [Figure 4C]. In addition, La promotes carbon nanotube formation instead of parasitic amorphous carbon coating to ensure nickel active sites are able to interact with the reactants and prolonged catalytic lifetime, as illustrated in Figure 4D.
Anionic substituted catalysts
The HAP surface property can be tuned via anion substitution at the B-site by substituting the phosphate group (PO4) with the vanadate group (VO4). Calcium vanadate apatite [Ca10(PO4)6-x(VO4)x(OH)2)] was first prepared using coprecipitation of calcium and vanadate aqueous solution with ammonium phosphate dibasic for oxidative PDH[68].
The addition of vanadium to the HAP catalyst induces the formation of lattice oxygen during the reaction due to the structure of the vanadate group. Sugiyama et al. reported that the presence of a surface hydroxyl group is beneficial for propane conversion as compared to one without a surface hydroxyl group and that was observed on a partially substituted vanadate HAP catalyst[68]. Partial substitution of the phosphate group with the vanadate group enables the presence of a surface hydroxyl group which maintains the acid-base properties and enables propane activation[68]. Lattice oxygen abstraction creates a redox cycle whereby the vanadium valence charge is reduced from V+5 to V+4[151]. This redox behavior is an important characteristic of the oxidative dehydrogenation of propane to propene. Recently, Petit et al. have tried to explain this phenomenon by employing in-situ DRIFTS and operando ionic conductance characterization comprehensively[69]. The partial substitution of the vanadate group into the phosphate group creates more defects and results in fewer OH groups. This causes charge balance on the HAP surface, which results in the creation of O2- species on the surface at a high temperature (723 K) that was shown to be active in propane activation and conversion to propene, as shown in the proposed mechanism in Figure 5A. Hydroxyl groups in the channels may play a part in activating propane molecules through the creation of O2- groups. The anion exchange between the PO43- and CO32- anions may also occur through the ball-milling method, which can increase oxygen vacancy concentration on the PO43- site[152].
Figure 5. (A) Schematized representation of the catalytic transformation of propane to propylene via non-oxidative and oxidative dehydrogenation routes on V-HAP samples with Ca2+-O2- surface acid-base pairs generated due to proton migration process close to VO4 tetrahedra. The square box represents oxygen vacancy and the squiggly arrow refers to the proton migration process. Reproduced from Ref.[69] with permission from Wiley-VCH GmbH. (B) Product selectivity for OCM reactions over HAP-based catalysts at 23% conversion under 973 K and 101 kPa pressure conditions and a space velocity of 8,800 mL·gcat-1·h-1. Reproduced from Ref.[58] with permission from Elsevier. (C) c-surface HAP (i), a-surface HAP (ii), the CH4 conversion rate of the Pb-HAP catalyst with different orientations for OCM reaction. Reproduced from Ref.[60] with permission from American Chemical Society.
The HAP framework is also flexible towards chlorine storage via anion exchange with the surface hydroxyl groups, also known as A-site substitution, to form chlorapatite. The introduction of chlorine into the HAP framework can be done using tetrachloromethane (TCM). Pre-treatment of the Sr-HAP catalyst with TCM causes the formation of SrCl2 and better ethylene yield can be observed as complete transformation to
Combining cationic and anionic substitution can be done with HAP. Oh et al. exploited the tunable nature of the cationic and anionic sites in the HAP framework by using the Pb2+ cation and CO32- anion to form Pb-HAP-CO3[58]. The product selectivity and CH4 conversion were tuned as described in Figure 5B as Pb2+ ions can stabilize methyl radicals and facilitate the coupling of methyl radicals to form C2 by forming covalent bonds with carbon[59]. The presence of the CO32- group appeared to maintain the HAP structure under harsh OCM conditions. However, the role of the CO32- group in stabilizing the HAP structure remains unclear. As different crystal facets can affect the catalytic activity and product selectivity, Oh et al. synthesized two types of HAP catalyst support: one with basal-faceted plane c-surface and prism-faceted a-surface as shown in Figure 5C(i-ii)[60]. The HAP and Pb-HAP with a c-surface performed better than the ones with the a-surface as oxide vacancies and OH- vacancies found on the c-surface facilitate the formation of methyl or methylene group, which are important to ethylene or ethane from the radical coupling [Figure 5C(iii)]. Thus, there is room for improvement in catalytic activity through a better understanding of the structure-activity relationship of the HAP catalyst through operando or in-situ characterization of the HAP surface structure.
Use of promoters in HAP catalysts
Alkali, alkaline earth, and noble metal promoters are frequently used in heterogeneous catalysis for positive benefits in catalytic activity and selectivity. In general, there are two types of promoters: structural and textural. Structural promoters enhance the desired product selectivity for a certain reaction by making the desired reaction pathway more energetically favorable[153]. Textural promoters attenuate catalyst nanoparticle agglomeration due to metal sintering; thereby, increasing catalytic stability[154,155].
In several reports on CO oxidation, the active site is frequently poisoned by the formation of carbonate, which leads to reduced catalytic activity. Basic elements such as Na can be added to the catalyst to decrease the number of basic sites, enabling better desorption of CO2 and resulting in less poisoning of the active sites[156]. Noble metal promoters are used for CO oxidation. Other than Au, Ag is also used as a cheaper alternative for CO oxidation in recent years. Silver doping in the HAP framework produced stable and active CO oxidation catalyst at high temperatures. The silver-doped HAP catalyst also demonstrated stable catalytic performance despite multiple cycles. However, the silver-doped HAP catalyst shows favorable room for improvement[157].
For the WGS reaction, the increase in surface basicity by the addition of alkaline earth or basic elements such as Sr, Na, and K helps to strengthen CO adsorption and favors water dissociation[158-160]. Iriarte et al. used calcined pork bone with HAP as the main phase and a minute amount of potassium as the support to be impregnated with Ni, Co, Cu and Fe[161]. The Ni-impregnated catalyst is found to be more active and selective to H2 formation than the other catalyst due to the presence of trace amounts of potassium, which suppresses methane formation. K is postulated to block the active site for CO methanation reaction. This study shows that a minute amount of alkaline promoter in naturally sourced material such as bones can be beneficial towards WGS reaction. It also highlights bones as a rich and accessible source of HAP when properly treated.
The addition of Lanthanum Oxide (La2O3) as a promoter has been explored in HAP. The increased La loading at fixed Ni loading enables higher CO2 conversion and CH4 selectivity in CO2 methanation, as shown in Figure 6A and B[135]. The increased addition of La2O3 introduces strong basic sites at greater concentrations, helping to strengthen CO2 chemisorption to the catalyst surface, which aids in CO2 conversion and eventual CH4 formation, as observed in Figure 6C. Additionally, La2O3 can enhance the dispersion of Ni nanoparticles, thereby making the catalyst more reducible. Other metals can also enhance the dispersion of Ni nanoparticles, such as alloying strategy with Co[162] or Ca as a monolayer[163]. In addition, the Co was demonstrated to inhibit graphitic carbon formation in DRM, thereby improving catalytic stability.
Figure 6. (A) CO2 conversion and (B) CH4 selectivity of the Ni/La-(x) catalysts in the CO2 methanation reaction. Reaction mixture conditions: 16% CO2 and 64% H2, balanced in He (WHSV = 30,000 mL/g.h). Reproduced from Ref.[135] with permission from Elsevier. (C) CO2-TPD (Left) and CH4 formation during TPSR experiments (Right) with pre-adsorbed CO2 on the reduced Ni/La-(x) catalysts. Reproduced from Ref.[135] with permission from Elsevier. (D) Reusability studies of catalysts with leaching content of K+ ions. Reproduced from Ref.[75] with permission from Elsevier.
Ceria is usually added as a promoter to increase the presence of oxygen species. The presence of oxygen species from Ce helps to alleviate carbon formation by promoting the Bouduoard reaction
Other than CO oxidation and WGS reaction, the transesterification reaction requires a highly basic or alkaline element to accelerate the reaction towards FAME formation. To enhance the alkalinity of the HAP catalyst surface, K2CO3, Sr, and Li were impregnated onto the bone-derived HAP catalyst support, respectively. The FAME yield increases in the following order: Li/HAP (93.2%) < Sr/HAP (94.7%) < 30K/HAP-600 (96.4%)[75]. The better catalytic performance exhibited by the 30K/HAP-600 catalyst is due to the higher strength of the basic sites and greater concentration of surface basic sites. The FAME yield was found to be not just dependent on the K2CO3 loading but also on the calcination temperature. However, if the calcination temperature is too high, the FAME yield is observed to decrease, which may be attributed to
Single atom and cluster on HAP catalyst
Single-atom catalysts (SAC) have received extensive research attention due to their promise of high atom efficiency and favorable metal-support interaction[167]. Ion exchange method is most commonly used to synthesize SAC using HAP as catalyst support. The seminal work by Yamaguchi et al. used the ion-exchange method to produce single-atom Ru3+, which substitutes Ca2+ in the HAP framework[168]. The presence of Ru3+ had been evidenced using XANES and EXAFS, as observed in Figure 7A. The Ru3+ species are observed to be coordinated to the Cl atom, which produces a highly active catalyst for aerobic alcohol oxidation. The same ion exchange method has been demonstrated for single atomic Pd[143], and Co[169], on HAP. Theoretical and computational study was conducted on single atomic transition metals (i.e., Fe, Co, Ni, and Cu) on HAP, revealing their suitability in direct POM to methanol[170]. The ion-exchange method can also be used to deposit clusters on HAP support as well. Tounsi et al. reported that the time taken for ion exchange to occur plays an important role in influencing the nature of Cu species and Cu loading as Cu2+ species will first chemisorb onto the HAP surface, followed by a slower reaction to be integrated into the HAP framework[171]. Although Tounsi et al. claimed that CuO was present on the HAP, no TEM and XANES characterization were shown to support this claim[171]. Thus, the effect of ion-exchange time on cluster formation requires further investigation. Lu et al. performed DFT on Fe-exchanged HAP for oxygen reduction reaction (ORR)[172]. The first principles study indicated that Fe species exchanged at Ca(II) site is more favorable. The Fe species at that site is favorable towards oxygen chemisorption which is the highest energy barrier to overcome to start ORR. The electrostatic adsorption method was used to create highly dispersed Ni-active sites on Ce-doped HAP [Figure 7B], which was characterized by higher activity and stability in DRM[89]. Although active sites in the form of single atoms and clusters can be deposited onto HAP support, the mesoporosity of the HAP support can be increased to enhance active site dispersion.
Figure 7. (A) FT magnitude of k3-weighted EXAFS of RuHAP with an inset showing a proposed surface around Ru3+ of the RuHAP catalyst (Left) and inverse FT of the peaks with the 0.8 < R/Å < 2.8 range (Right). Adapted from Ref.[168] with permission from American Chemical Society. (B) Electron microscopy images and size distribution with evidence of single-atom Ni in 0.5Ni1/HAP-Ce samples. Adapted from Ref.[89] under CC BY 4.0 license.
Mesoporosity in HAP-based catalyst
The ability to control the catalyst pore size can have a profound effect on product selectivity[173]. The catalyst pore size can be tuned by two methods: soft-template and hard-template. The effect of pore structure in the HAP support on the DRM performance of Ni/HAP catalysts was reported by Li et al.[97]. The mesopore/macropore ratio of the support can be altered by adjusting the urea-hydrolysis precipitation temperature. This ratio dictates Ni dispersion and the size and interaction of NiO nanoparticles with the support. A high fraction of macropores results in poor Ni dispersion and low metal-support interaction. As mesopores increase, Ni preferentially disperses along mesopore channel walls, resulting in smaller nanoparticles with enhanced metal-support interaction and improved activity. However, too small pore sizes cause blockage by Ni nanoparticles, leading to suboptimal DRM activity. In another study, the specific surface area and pore structure of the HAP support were modified using various surfactants in a soft-templating synthesis approach. Soft templating by a combination of two surfactants was reported to significantly increase the surface area and enhance reactant conversion in DRM[146]. The use of surfactants such as CTAB during the synthesis stage can help to increase BET surface area. Essamlali et al. utilized CTAB to increase the synthetic HAP surface area, in addition, to increasing surface basicity by impregnating with Na[79]. The hard templating method first involves the hydrothermal treatment of SBA-15 in sucrose solution to produce carbonized sucrose in the SBA-15 pores[174]. The SBA-15 SiO2 hard template was removed by washing with concentrated NaOH to obtain carbon nanorods. Those carbon nanorods were added to the coprecipitation solution for HAP nanoparticle synthesis. The BET surface area of the resultant HAP after air calcination is 242.2 m2/gcat, which is among the highest BET surface areas reported for the HAP material.
The HAP BET surface area can be increased using various treatment methods. Bone-derived HAP catalyst underwent two preparation methods: (1) hydrothermal treatment after calcination, and (2) without hydrothermal treatment after calcination[175]. The hydrothermal-treated sample showed a higher BET surface area and FAME yield, indicating improved access to active sites due to increased surface area and mesopores, as shown in the BET results. However, no relationship was found between hydrothermal duration and catalytic activity. Ghanei et al. used waste bone and reported low catalytic activity[78]. Their procedure indicated an average particle size of 180-300 μm, resulting in a lower BET surface area of
FUTURE RESEARCH DIRECTIONS AND OUTLOOK
The HAP unit cell structure accommodates various cations due to similar ionic radii and valence charges. Cations integrate into HAP through surface complexation, which involves liquid-solid mass transfer that can create resistance. Therefore, it is crucial to explore strategies to speed up the substitution. The Ca/P ratio also significantly affects the HAP surface basicity or acidity. For DRM, the presence of basic sites generally helps to alleviate carbon filament formation, thereby extending the longevity of the catalytic performance. For other catalytic reactions such as WGS reaction, higher surface basicity helps with enhancing water dissociation; thereby, higher H2 formation[160]. To better understand the surfaces of HAP-based catalyst surfaces, in-situ or operando spectroscopic techniques such as DRIFTS, XPS, and X-ray absorption spectroscopy (XAS) can be used to gain insight into the active metal (e.g., Ni)-Ca coordination, which may be important to understand metal-support interaction.
According to the literature examined in the review manuscript, Ca(II) is observed to be the site of choice for cation substitution. Limited studies attempted to investigate the concentration of Ni2+ substituted into the Ca(I) or Ca(II) sites[90]. However, the structure-activity relationship between DRM activity and the concentration of Ni2+ substituted into the Ca(I) or Ca(II) remains unclear. Using DFT and microkinetic modeling to clarify the structure-activity relationship for DRM would be interesting, involving examining how catalytic activity improves with varying Ni2+ concentrations in Ca(I) or Ca(II) sites. In fact, DFT studies have been employed in other reaction studies, such as the ORR using Fe-doped HAP, where DFT uncovered superiority in reducing O2 to H2O with less kinetic and thermodynamic barrier in the OH hydrogenation or 2OH self-dehydration in Fe-substituted HAP (i.e., Fe@HAP) compared to Fe/HAp[172]. DFT and microkinetic studies were employed to uncover how the unique combination of acidic and basic properties of HAP facilitates different elementary reactions in ethanol-to-butanol biofuels production[176], or how defects in HAP can stabilize intermediate reaction species[177]. Importantly, DFT is not only a tool for understanding cationic substitutability as earlier elucidated[42,43], but can reveal key mechanistic insights not limited to preferential reaction pathways or adsorption affinity on different active sites (i.e., active metals, or defects on HAP)[132,178]. Substitution with foreign anion at either the OH- group or PO42- group can affect surface properties of the HAP surface, especially surface acidity and basicity. Currently, there is a scarcity of literature on HAP-based catalysts examining the effect of anionic substitution on catalytic activity, except for the ones on oxidative PDH and OCM. More experimental studies on the effect of anionic substitution on catalytic activity and product selectivity for CO2 methanation, WGS reaction, CO oxidation, and DRM can be examined in greater detail. For example, studying how CO2 can be absorbed in HAP through anionic substitution as carbonated HAP[179,180] or possibly with exsolved calcium ions after cationic substitution may be an interesting avenue of research for identifying efficient, low-cost carbon capture or chemical looping materials[181].
Core-shell nanoarchitectures involving the HAP nanoparticles were primarily used for gas sensors and non-catalytic applications[182-184]. HAP-based catalysts have shown promise in high-temperature reactions such as DRM and OCM. Their catalytic stability can be improved by encapsulating HAP nanoparticles with inert shells (e.g., SiO2) using methods such as Stöber synthesis or reverse microemulsion. Promoters play a critical role by blocking undesired reaction pathways, enhancing reactant adsorption, and improving product selectivity. Alkaline or basic promoters facilitate water dissociation, aiding WGS reaction rates, and also enhance the reducibility of active metal centers.
In CO2 capture and utilization, such promoters enable dual-function materials that both adsorb and convert CO2[185]. During flue gas exposure, CO2 is captured at basic sites; when switched to H2, the adsorbed CO2 is converted into products such as CH3OH or CH4 at metal sites. These promoters also increase surface basicity, which improves biodiesel yields. To fully understand their role, it is important to examine whether they also influence catalyst structure and texture, using both ex-situ and in-situ characterizations on spent catalysts.
Literature reports that the BET surface area of HAP-based catalysts can vary widely, from as low as 5 up to 242.2 m2/g. The highest values are typically achieved using hard templating methods, such as SBA-15-derived carbon nanorods. However, this approach is complex and time-consuming. Despite the promise of mesoporous HAP, its actual use in catalysis remains limited. Alternatively, poly(methyl methacrylate) (PMMA) can serve as a hard template that burns off easily, potentially saving synthesis time and reducing costs. Although natural bones can still be used as a source of HAP, the BET surface area is still very low
More recently, piezoelectric catalysis has gained increasing research attention to generate reactive oxygen species (ROS) - hydroxyl radicals (OH·) can activate C-H bond in methane at milder conditions to facilitate C-C coupling to form hydrocarbon compounds and oxygenates[186,187]. Piezoelectric materials, upon application of mechanical stress, can induce the asymmetrical distribution of surface charge and start an electrochemical reaction[188]. HAP is found to possess piezoelectric properties, which can be exploited to generate ROS[189,190]. Zhou et al. recently studied the use of HAP as a potential material for piezoelectric methane oxidation[191]. Interestingly, the HAP material, upon ultrasonic-induced stresses, produces hydroxyl radical (•OH) to attack the methane C-H bond and negative surface charges to drive methanol dehydration into methyl carbene as a C1 precursor for upgrading into C2-C3 alcohols, as shown in Figure 8. Further examination of the surface engineering strategies discussed in this review on HAP as piezoelectric catalysts could be an interesting future research focus.
We have covered sustainable chemical production via OCM, PDH, biodiesel production, and H2 production and/or CO2-modulating reactions, including CO2-utilizing DRM, POM, CO oxidation, and CO2 methanation. However, other reactions relevant to sustainable chemical and fuel production include CO2 hydrogenation to methanol, CO2 hydrogenation to olefins and aromatics, and the upgrading of alcohol (ethanol) to higher alcohols[192], which can benefit from HAP-based catalysts. However, these are explored only to a limited extent. Additionally, to the best of our knowledge, there seems to be a scarcity of literature regarding the economics of HAP-based catalyst production. According to Guo et al., the cost to produce the copper HAP catalyst is approximately USD 1.0 per kilogram[133]. However, this calculation does not account for key factors such as operating costs and bulk raw material pricing. HAP is often viewed as a green and sustainable catalyst, especially because it can be derived from bio-based waste sources such as animal bones and fish scales. This not only reduces environmental impact but also improves the overall sustainability of HAP-based catalytic systems. That said, the feasibility of scaling up HAP - whether to pilot[193] or commercial levels - still requires further investigation, particularly in terms of supply chain and catalytic performance.
CONCLUSION
The propensity of HAP material for cation and/or anion substitution enhances surface basicity, creates active sites, and introduces lattice oxygen species. As a result, new active sites can be formed for catalytic reactions that are crucial for chemical and fuel products such as DRM, CO2 hydrogenation, POM, CO oxidation, and biodiesel production. The basicity of the HAP surface can be adjusted by adding basic and alkaline elements such as K, Na, and Sr. Incorporating these elements helps improve product selectivity and reactant conversion. For certain reactions that rely on precious metals as active sites, metal loading can be reduced by employing single-atom or cluster catalysts. Cationic substitution in HAP can be utilized to generate such single atoms and clusters, which may be catalytically active, selective, and stable for reactions essential to developing more sustainable fuels and chemical products such as CO2 hydrogenation and DRM. Enhancements in these catalytic properties would lead to higher product yields, thereby lowering manufacturing costs. Improvements in catalytic performance using HAP-based catalysts can be achieved through a deeper understanding of the HAP surface, utilizing synergistic surface characterization techniques such as XAS, XPS, and FTIR in either in-situ or operando mode.
DECLARATIONS
Authors' contributions
Wrote the manuscript: Lim, K. H.; Wai, M. H.
Reviewed the manuscript: Cao, K.; Das, S.; Nzihou, A.; Kawi, S.
Availability of data and materials
Not applicable.
Financial support and sponsorship
The authors acknowledge the financial support by the National Research Foundation, Singapore, and A*STAR under its Low-Carbon Energy Research (LCER) Funding Initiative (FI) Project (U2102d2011, WBS: A-8000278-00-00).
Conflicts of interest
Kawi, S. is an Editorial Board Member of the journal Energy Materials but is not involved in any steps of editorial processing, notably including reviewer selection, manuscript handling, or decision-making, while the other authors have declared that they have no conflicts of interest.
Ethical approval and consent to participate
Not applicable.
Consent for publication
Not applicable.
Copyright
© The Author(s) 2025.
REFERENCES
1. Lim, K. H.; Yue, Y.; Bella,
2. Julkapli, N. M.; Bagheri, S. Graphene supported heterogeneous catalysts: an overview. Int. J. Hydrogen. Energy. 2015, 40, 948-79.
4. Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T. Metal-organic framework materials as catalysts. Chem. Soc. Rev. 2009, 38, 1450-9.
5. Jiao, L.; Wang, Y.; Jiang, H. L.; Xu, Q. Metal-organic frameworks as platforms for catalytic applications. Adv. Mater. 2018, 30, e1703663.
6. Wang, Q.; Astruc, D. State of the art and prospects in metal-organic framework (MOF)-based and MOF-derived nanocatalysis. Chem. Rev. 2020, 120, 1438-511.
7. Santos, V. P.; Wezendonk, T. A.; Jaén, J. J.; et al. Metal organic framework-mediated synthesis of highly active and stable Fischer-Tropsch catalysts. Nat. Commun. 2015, 6, 6451.
8. Frei, M. S.; Mondelli, C.; García-Muelas, R.; et al. Atomic-scale engineering of indium oxide promotion by palladium for methanol production via CO2 hydrogenation. Nat. Commun. 2019, 10, 3377.
9. Wang, J.; Li, W. C.; Sun, D. H.; He, L.; Zhou, B. C.; Lu, A. H. High-selective upgrading of ethanol to C4-10 alcohols over hydroxyapatite catalyst with superior basicity. ACS. Sustain. Chem. Eng. 2025, 13, 36-45.
10. Tsuchida, T.; Sakuma, S.; Takeguchi, T.; Ueda, W. Direct synthesis of n-butanol from ethanol over nonstoichiometric hydroxyapatite. Ind. Eng. Chem. Res. 2006, 45, 8634-42.
11. Qi, H.; Li, Y.; Zhou, Z.; et al. Synthesis of piperidines and pyridine from furfural over a surface single-atom alloy Ru1CoNP catalyst. Nat. Commun. 2023, 14, 6329.
12. Ibrahim, M.; Labaki, M.; Giraudon, J. M.; Lamonier, J. F. Hydroxyapatite, a multifunctional material for air, water and soil pollution control: a review. J. Hazard. Mater. 2020, 383, 121139.
13. Kaneda, K.; Mizugaki, T. Development of concerto metal catalysts using apatite compounds for green organic syntheses. Energy. Environ. Sci. 2009, 2, 655.
14. Fihri, A.; Len, C.; Varma, R. S.; Solhy, A. Hydroxyapatite: a review of syntheses, structure and applications in heterogeneous catalysis. Coordin. Chem. Rev. 2017, 347, 48-76.
15. Sadat-Shojai, M.; Khorasani, M. T.; Dinpanah-Khoshdargi, E.; Jamshidi, A. Synthesis methods for nanosized hydroxyapatite with diverse structures. Acta. Biomater. 2013, 9, 7591-621.
16. Pham Minh, D. Introduction to hydroxyapatite-based materials in heterogeneous catalysis; 2022; pp. 1-18.
17. Mostafa, N. Y.; Brown, P. W. Computer simulation of stoichiometric hydroxyapatite: structure and substitutions. J. Phys. Chem. Solids. 2007, 68, 431-7.
18. Ammar, M.; Ashraf, S.; Baltrusaitis, J. Nutrient-doped hydroxyapatite: structure, synthesis and properties. Ceramics 2023, 6, 1799-825.
19. Tsuchida, T.; Kubo, J.; Yoshioka, T.; Sakuma, S.; Takeguchi, T.; Ueda, W. Reaction of ethanol over hydroxyapatite affected by Ca/P ratio of catalyst. J. Catal. 2008, 259, 183-9.
20. Bailliez, S.; Nzihou, A.; Bèche, E.; Flamant, G. Removal of lead (Pb) by hydroxyapatite sorbent. Process. Saf. Environ. Prot. 2004, 82, 175-80.
21. Pizzala, H.; Caldarelli, S.; Eon, J. G.; et al. A solid-state NMR study of lead and vanadium substitution into hydroxyapatite. J. Am. Chem. Soc. 2009, 131, 5145-52.
22. Robles-Águila, M.; Reyes-Avendaño, J.; Mendoza, M. Structural analysis of metal-doped (Mn, Fe, Co, Ni, Cu, Zn) calcium hydroxyapatite synthetized by a sol-gel microwave-assisted method. Ceram. Int. 2017, 43, 12705-9.
23. Takahashi, H.; Yashima, M.; Kakihana, M.; Yoshimura, M. A differential scanning calorimeter study of the monoclinic (P21/b)↔hexagonal (P63/m) reversible phase transition in hydroxyapatite. Thermochim. Acta. 2001, 371, 53-6.
24. Ma, G.; Liu, X. Y. Hydroxyapatite: hexagonal or monoclinic? Cryst. Growth. Des. 2009, 9, 2991-4.
25. Van Rees, H. B.; Mengeot, M.; Kostiner, E. Monoclinic-hexagonal transition in hydroxyapatite and deuterohydroxyapatite single crystals. Mater. Res. Bull. 1973, 8, 1307-9.
26. Li, W.; Zhang, G.; Jiang, X.; et al. CO2 hydrogenation on unpromoted and M-promoted Co/TiO2 catalysts (M = Zr, K, Cs): effects of crystal phase of supports and metal-support interaction on tuning product distribution. ACS. Catal. 2019, 9, 2739-51.
27. Jiang, F.; Wang, S.; Liu, B.; et al. Insights into the influence of CeO2 crystal facet on CO2 hydrogenation to methanol over Pd/CeO2 catalysts. ACS. Catal. 2020, 10, 11493-509.
28. Zhao, Z.; Jiang, Q.; Wang, Q.; et al. Effect of rutile content on the catalytic performance of Ru/TiO2 catalyst for low-temperature CO2 methanation. ACS. Sustain. Chem. Eng. 2021, 9, 14288-96.
29. Moteki, T.; Flaherty, D. W. Mechanistic insight to C-C bond formation and predictive models for cascade reactions among alcohols on Ca- and Sr-hydroxyapatites. ACS. Catal. 2016, 6, 4170-83.
30. Bett, J. A. S.; Christner, L. G.; Hall, W. K. Hydrogen held by solids. XII. Hydroxyapatite catalysts. J. Am. Chem. Soc. 1967, 89, 5535-41.
31. Wang, Y. B.; He, L.; Zhou, B. C.; et al. Hydroxyapatite nanorods rich in [Ca-O-P] sites stabilized Ni species for methane dry reforming. Ind. Eng. Chem. Res. 2021, 60, 15064-73.
32. Lin, T.; Liu, C.; Gangarajula, Y.; et al. Strong metal-support interaction induced sintering-resistant Ni nanoparticles supported on highly CO2-activating vanadium hydroxyapatite for dry reforming of methane. Appl. Cata. A. Gen. 2023, 662, 119290.
33. Ferri, M.; Campisi, S.; Scavini, M.; Evangelisti, C.; Carniti, P.; Gervasini, A. In-depth study of the mechanism of heavy metal trapping on the surface of hydroxyapatite. Appl. Surf. Sci. 2019, 475, 397-409.
34. Jemli, Y. E. L.; Abdelouahdi, K.; Minh, D. P.; Barakat, A.; Solhy, A. Synthesis and characterization of hydroxyapatite and hydroxyapatite-based catalysts. In: Pham Minh D, editor. Design and applications of hydroxyapatite-based catalysts. Wiley; 2022. pp. 19-72.
35. Smiciklas, I.; Onjia, A.; Raicević, S.; Janaćković, D.; Mitrić, M. Factors influencing the removal of divalent cations by hydroxyapatite. J. Hazard. Mater. 2008, 152, 876-84.
36. Shimabayashi, S.; Tamura, C.; Nakagaki, M. Adsorption of mono- and divalent metal cations on hydroxyapatite in water. Chem. Pharm. Bull. 1981, 29, 2116-22.
37. Takeuchi, Y.; Arai, H. Removal of coexisting Pb2+, Cu2+ and Cd2+ ions from water by addition of hydroxyapatite powder. J. Chem. Eng. Japan. 1990, 23, 75-80.
38. Peng, S.; Lin, Y.; Lee, W.; Lin, Y.; Hung, M.; Lin, K. Removal of Cu2+ from wastewater using eco-hydroxyapatite synthesized from marble sludge. Mater. Chem. Phys. 2023, 293, 126854.
39. Geng, Z.; Cui, Z.; Li, Z.; et al. Synthesis, characterization and the formation mechanism of magnesium- and strontium-substituted hydroxyapatite. J. Mater. Chem. B. 2015, 3, 3738-46.
40. Saito, T.; Yokoi, T.; Nakamura, A.; Matsunaga, K. Formation energies and site preference of substitutional divalent cations in carbonated apatite. J. Am. Ceram. Soc. 2020, 103, 5354-64.
41. Oh, S. C.; Lei, Y.; Chen, H.; Liu, D. Catalytic consequences of cation and anion substitutions on rate and mechanism of oxidative coupling of methane over hydroxyapatite catalysts. Fuel 2017, 191, 472-85.
42. Liu, H.; Cui, X.; Lu, X.; Liu, X.; Zhang, L.; Chan, T. Mechanism of Mn incorporation into hydroxyapatite: Insights from SR-XRD, Raman, XAS, and DFT calculation. Chem. Geol. 2021, 579, 120354.
43. Ellis, D. E.; Terra, J.; Warschkow, O.; et al. A theoretical and experimental study of lead substitution in calcium hydroxyapatite. Phys. Chem. Chem. Phys. 2006, 8, 967-76.
44. Kandori, K.; Toshima, S.; Wakamura, M.; Fukusumi, M.; Morisada, Y. Effects of modification of calcium hydroxyapatites by trivalent metal ions on the protein adsorption behavior. J. Phys. Chem. B. 2010, 114, 2399-404.
45. Singh, G.; Jolly, S. S.; Singh, R. P. Cerium substituted hydroxyapatite mesoporous nanorods: synthesis and characterization for drug delivery applications. Mater. Today. Proc. 2020, 28, 1460-6.
46. Ciobanu, G.; Maria, Bargan., A.; Luca, C. New cerium(IV)-substituted hydroxyapatite nanoparticles: preparation and characterization. Ceram. Int. 2015, 41, 12192-201.
47. Ivanova, T.; Frank-Kamenetskaya, O.; Kol'tsov, A.; Ugolkov, V. Crystal structure of calcium-deficient carbonated hydroxyapatite. thermal decomposition. J. Solid. State. Chem. 2001, 160, 340-9.
48. Resende, N. S.; Nele, M.; Salim, V. M. Effects of anion substitution on the acid properties of hydroxyapatite. Thermochim. Acta. 2006, 451, 16-21.
49. Astala, R.; Stott, M. J. First principles investigation of mineral component of bone: CO3 substitutions in hydroxyapatite. Chem. Mater. 2005, 17, 4125-33.
50. Lunsford, J. H. The catalytic oxidative coupling of methane. Angew. Chem. Int. Ed. 1995, 34, 970-80.
51. Arndt, S.; Otremba, T.; Simon, U.; Yildiz, M.; Schubert, H.; Schomäcker, R. Mn-Na2WO4/SiO2 as catalyst for the oxidative coupling of methane. What is really known? Appl. Catal. A. Gen. 2012, 425-6, 53-61.
52. Feng, R.; Niu, P.; Wang, Q.; et al. In-depth understanding of the crystal-facet effect of La2O2CO3 for low-temperature oxidative coupling of methane. Fuel 2022, 308, 121848.
53. Wang, Z.; Wang, D.; Gong, X. Strategies to improve the activity while maintaining the selectivity of oxidative coupling of methane at La2O3: a density functional theory study. ACS. Catal. 2020, 10, 586-94.
54. Park, J. H.; Lee, D.; Im, S.; et al. Oxidative coupling of methane using non-stoichiometric lead hydroxyapatite catalyst mixtures. Fuel 2012, 94, 433-9.
55. Lee, K. Y.; Han, Y. C.; Suh, D. J.; Park, T. J. Pb-substituted hydroxyapatite catalysts prepared by coprecipitation method for oxidative coupling of methane. Stud. Surf. Sci. Catal. 1998, 119, 385-90.
56. Bae, Y. K.; Jun, J. H.; Yoon, K. J. Oxidative coupling of methane over promoted strontium chlorapatite. Korean. J. Chem. Eng. 1999, 16, 595-601.
57. Sugiyama, S.; Fujii, Y.; Hayashi, H. Oxidative coupling of methane on calcium-lead and barium-lead hydroxyapatites. Phosphorus. Res. Bull. 2002, 14, 111-8.
58. Oh, S. C.; Wu, Y.; Tran, D. T.; Lee, I. C.; Lei, Y.; Liu, D. Influences of cation and anion substitutions on oxidative coupling of methane over hydroxyapatite catalysts. Fuel 2016, 167, 208-17.
59. Matsumura, Y.; Sugiyama, S.; Hayashi, H.; Moffat, J. B. Lead-calcium hydroxyapatite: cation effects in the oxidative coupling of methane. J. Solid. State. Chem. 1995, 114, 138-45.
60. Oh, S. C.; Xu, J.; Tran, D. T.; Liu, B.; Liu, D. Effects of controlled crystalline surface of hydroxyapatite on methane oxidation reactions. ACS. Catal. 2018, 8, 4493-507.
61. Xiang, D.; Li, P.; Yuan, X. Process optimization, exergy efficiency, and life cycle energy consumption-GHG emissions of the propane-to-propylene with/without hydrogen production process. J. Clean. Prod. 2022, 367, 133024.
62. Atanga, M. A.; Rezaei, F.; Jawad, A.; Fitch, M.; Rownaghi, A. A. Oxidative dehydrogenation of propane to propylene with carbon dioxide. Appl. Catal. B. Environ. 2018, 220, 429-45.
63. Carter, J. H.; Bere, T.; Pitchers, J. R.; et al. Direct and oxidative dehydrogenation of propane: from catalyst design to industrial application. Green. Chem. 2021, 23, 9747-99.
64. Yu, C.; Ge, Q.; Xu, H.; Li, W. Influence of oxygen addition on the reaction of propane catalytic dehydrogenation to propylene over modified Pt-based catalysts. Ind. Eng. Chem. Res. 2007, 46, 8722-8.
65. Barman, S.; Maity, N.; Bhatte, K.; et al. Single-site VOx moieties generated on silica by surface organometallic chemistry: a way to enhance the catalytic activity in the oxidative dehydrogenation of propane. ACS. Catal. 2016, 6, 5908-21.
66. Sugiyama, S.; Shono, T.; Makino, D.; Moriga, T.; Hayashi, H. Enhancement of the catalytic activities in propane oxidation and H-D exchangeability of hydroxyl groups by the incorporation with cobalt into strontium hydroxyapatite. J. Catal. 2003, 214, 8-14.
67. Boucetta, C.; Kacimi, M.; Ensuque, A.; Piquemal, J.; Bozon-Verduraz, F.; Ziyad, M. Oxidative dehydrogenation of propane over chromium-loaded calcium-hydroxyapatite. Appl. Catal. A. Gen. 2009, 356, 201-10.
68. Sugiyama, S.; Osaka, T.; Hirata, Y.; Sotowa, K. Enhancement of the activity for oxidative dehydrogenation of propane on calcium hydroxyapatite substituted with vanadate. Appl. Catal. A. Gen. 2006, 312, 52-8.
69. Petit, S.; Thomas, C.; Millot, Y.; Krafft, J.; Laberty-Robert, C.; Costentin, G. Activation of C-H bond of propane by strong basic sites generated by bulk proton conduction on V-modified hydroxyapatites for the formation of propene. ChemCatChem 2020, 12, 2506-21.
70. Smith, S. M.; Oopathum, C.; Weeramongkhonlert, V.; et al. Transesterification of soybean oil using bovine bone waste as new catalyst. Bioresour. Technol. 2013, 143, 686-90.
71. Al-Sakkari, E. G.; Mohammed, M. G.; Elozeiri, A. A.; et al. Comparative technoeconomic analysis of using waste and virgin cooking oils for biodiesel production. Front. Energy. Res. 2020, 8, 583357.
72. Changmai, B.; Vanlalveni, C.; Ingle, A. P.; Bhagat, R.; Rokhum, S. L. Widely used catalysts in biodiesel production: a review. RSC. Adv. 2020, 10, 41625-79.
73. Badnore, A. U.; Jadhav, N. L.; Pinjari, D. V.; Pandit, A. B. Efficacy of newly developed nano-crystalline calcium oxide catalyst for biodiesel production. Chem. Eng. Process. Process. Intensif. 2018, 133, 312-9.
74. Marinković, D. M.; Stanković, M. V.; Veličković, A. V.; et al. Calcium oxide as a promising heterogeneous catalyst for biodiesel production: current state and perspectives. Renew. Sustain. Energy. Rev. 2016, 56, 1387-408.
75. Chen, G.; Shan, R.; Shi, J.; Liu, C.; Yan, B. Biodiesel production from palm oil using active and stable K doped hydroxyapatite catalysts. Energy. Convers. Manag. 2015, 98, 463-9.
76. Chakraborty, R.; Das, S. K. Optimization of biodiesel synthesis from waste frying soybean oil using fish scale-supported Ni catalyst. Ind. Eng. Chem. Res. 2012, 51, 8404-14.
77. Yan, B.; Zhang, Y.; Chen, G.; Shan, R.; Ma, W.; Liu, C. The utilization of hydroxyapatite-supported CaO-CeO2 catalyst for biodiesel production. Energy. Convers. Manag. 2016, 130, 156-64.
78. Ghanei, R.; Khalili Dermani, R.; Salehi, Y.; Mohammadi, M. Waste animal bone as support for CaO impregnation in catalytic biodiesel production from vegetable oil. Waste. Biomass. Valor. 2016, 7, 527-32.
79. Essamlali, Y.; Amadine, O.; Larzek, M.; Len, C.; Zahouily, M. Sodium modified hydroxyapatite: highly efficient and stable solid-base catalyst for biodiesel production. Energy. Convers. Manag. 2017, 149, 355-67.
80. Xie, W.; Han, Y.; Tai, S. Biodiesel production using biguanide-functionalized hydroxyapatite-encapsulated-γ-Fe2O3 nanoparticles. Fuel 2017, 210, 83-90.
81. Gupta, J.; Agarwal, M.; Dalai, A. K. Marble slurry derived hydroxyapatite as heterogeneous catalyst for biodiesel production from soybean oil. Can. J. Chem. Eng. 2018, 96, 1873-80.
82. Khan, H. M.; Iqbal, T.; Ali, C. H.; Yasin, S.; Jamil, F. Waste quail beaks as renewable source for synthesizing novel catalysts for biodiesel production. Renew. Energy. 2020, 154, 1035-43.
83. Alipour, Z.; Babu Borugadda, V.; Wang, H.; Dalai, A. K. Syngas production through dry reforming: a review on catalysts and their materials, preparation methods and reactor type. Chem. Eng. J. 2023, 452, 139416.
84. Göransson, K.; Söderlind, U.; He, J.; Zhang, W. Review of syngas production via biomass DFBGs. Renew. Sustain. Energy. Rev. 2011, 15, 482-92.
85. Fiore, M.; Magi, V.; Viggiano, A. Internal combustion engines powered by syngas: a review. Appl. Energy. 2020, 276, 115415.
86. Song, Y.; Ozdemir, E.; Ramesh, S.; et al. Dry reforming of methane by stable Ni-Mo nanocatalysts on single-crystalline MgO. Science 2020, 367, 777-81.
87. Zhou, Q.; Fu, X.; Hui Lim, K.; et al. Complete confinement of Ce/Ni within SiO2 nanotube with high oxygen vacancy concentration for CO2 methane reforming. Fuel 2022, 325, 124819.
88. Tanimu, A.; Yusuf, B. O.; Lateef, S.; et al. Dry reforming of methane: advances in coke mitigation strategies via siliceous catalyst formulations. J. Environ. Chem. Eng. 2024, 12, 113873.
89. Akri, M.; Zhao, S.; Li, X.; et al. Atomically dispersed nickel as coke-resistant active sites for methane dry reforming. Nat. Commun. 2019, 10, 5181.
90. Meng, J.; Pan, W.; Gu, T.; et al. One-pot synthesis of a highly active and stable Ni-embedded hydroxyapatite catalyst for syngas production via dry reforming of methane. Energy. Fuels. 2021, 35, 19568-80.
91. Rego de Vasconcelos, B.; Pham Minh, D.; Martins, E.; Germeau, A.; Sharrock, P.; Nzihou, A. Highly-efficient hydroxyapatite-supported nickel catalysts for dry reforming of methane. Int. J. Hydrogen. Energy. 2020, 45, 18502-18.
92. Tran, T. Q.; Pham Minh, D.; Phan, T. S.; Pham, Q. N.; Nguyen Xuan, H. Dry reforming of methane over calcium-deficient hydroxyapatite supported cobalt and nickel catalysts. Chem. Eng. Sci. 2020, 228, 115975.
93. Boukha, Z.; Kacimi, M.; Pereira, M. F. R.; Faria, J. L.; Figueiredo, J. L.; Ziyad, M. Methane dry reforming on Ni loaded hydroxyapatite and fluoroapatite. Appl. Catal. A. Gen. 2007, 317, 299-309.
94. Phan, T. S.; Sane, A. R.; Rêgo de Vasconcelos, B.; et al. Hydroxyapatite supported bimetallic cobalt and nickel catalysts for syngas production from dry reforming of methane. Appl. Catal. B. Environ. 2018, 224, 310-21.
95. Rêgo De Vasconcelos, B.; Zhao, L.; Sharrock, P.; Nzihou, A.; Pham Minh, D. Catalytic transformation of carbon dioxide and methane into syngas over ruthenium and platinum supported hydroxyapatites. App. Surf. Sci. 2016, 390, 141-56.
96. Akri, M.; El Kasmi, A.; Batiot-Dupeyrat, C.; Qiao, B. Highly active and carbon-resistant nickel single-atom catalysts for methane dry reforming. Catalysts 2020, 10, 630.
97. Li, B.; Yuan, X.; Li, B.; Wang, X. Impact of pore structure on hydroxyapatite supported nickel catalysts (Ni/HAP) for dry reforming of methane. Fuel. Proc. Technol. 2020, 202, 106359.
98. Rego de Vasconcelos, B.; Pham Minh, D.; Sharrock, P.; Nzihou, A. Regeneration study of Ni/hydroxyapatite spent catalyst from dry reforming. Catal. Today. 2018, 310, 107-15.
99. Dębek, R.; Motak, M.; Duraczyska, D.; et al. Methane dry reforming over hydrotalcite-derived Ni-Mg-Al mixed oxides: the influence of Ni content on catalytic activity, selectivity and stability. Catal. Sci. Technol. 2016, 6, 6705-15.
100. Parkinson, B.; Matthews, J. W.; Mcconnaughy, T. B.; Upham, D. C.; Mcfarland, E. W. Techno-economic analysis of methane pyrolysis in molten metals: decarbonizing natural gas. Chem. Eng. Technol. 2017, 40, 1022-30.
101. Elbadawi, A. H.; Ge, L.; Li, Z.; Liu, S.; Wang, S.; Zhu, Z. Catalytic partial oxidation of methane to syngas: review of perovskite catalysts and membrane reactors. Catal. Rev. 2021, 63, 1-67.
102. Jun, J. Nickel-calcium phosphate/hydroxyapatite catalysts for partial oxidation of methane to syngas: characterization and activation. J. Catal. 2004, 221, 178-90.
103. Matsumura, Y.; Moffat, J. Partial oxidation of methane to carbon-monoxide and hydrogen with molecular-oxygen and nitrous-oxide over hydroxyapatite catalysts. J. Catal. 1994, 148, 323-33.
104. Sugiyama, S.; Minami, T.; Higaki, T.; Hayashi, H.; Moffat, J. B. High selective conversion of methane to carbon monoxide and the effects of chlorine additives in the gas and solid phases on the oxidation of methane on strontium hydroxyapatites. Ind. Eng. Chem. Res. 1997, 36, 328-34.
105. Sugiyama, S.; Abe, K.; Minami, T.; Hayashi, H.; Moffat, J. B. A comparative study of the oxidation of methane and ethane on calcium hydroxyapatites with incorporated lead in the presence and absence of tetrachloromethane. Appl. Catal. A. Gen. 1998, 169, 77-86.
106. Sugiyama, S.; Fujii, Y.; Abe, K.; Hayashi, H.; Moffat, J. B. Facile formation of the partial oxidation and oxidative-coupling products from the oxidation of methane on barium hydroxyapatites with tetrachloromethane. Energy. Fuels. 1999, 13, 637-40.
107. Sugiyama, S. Oxidation of methane with nitrous oxide on calcium hydroxyapatites in the presence and absence of tetrachloromethane. J. Mol. Catal. A. Chem. 1999, 144, 347-55.
108. Jun, J. H.; Lee, S. J.; Lee, S.; et al. Characterization of a nickel-strontium phosphate catalyst for partial oxidation of methane. Korean. J. Chem. Eng. 2003, 20, 829-34.
109. Jun, J. H.; Jeong, K. S.; Lee, T.; et al. Nickel-calcium phosphate/hydroxyapatite catalysts for partial oxidation of methane to syngas: effect of composition. Korean. J. Chem. Eng. 2004, 21, 140-6.
110. Kim, K. H.; Lee, S. Y.; Nam, S.; Lim, T. H.; Hong, S.; Yoon, K. J. Promotion effects of ceria in partial oxidation of methane over Ni-calcium hydroxyapatite. Korean. J. Chem. Eng. 2006, 23, 17-20.
111. Jun, J.; Lim, T.; Nam, S.; Hong, S.; Yoon, K. Mechanism of partial oxidation of methane over a nickel-calcium hydroxyapatite catalyst. Appl. Catal. A. Gen. 2006, 312, 27-34.
112. Boukha, Z.; Gil-Calvo, M.; de Rivas, B.; González-Velasco, J. R.; Gutiérrez-Ortiz, J. I.; López-Fonseca, R. Behaviour of Rh supported on hydroxyapatite catalysts in partial oxidation and steam reforming of methane: on the role of the speciation of the Rh particles. Appl. Catal. A. Gen. 2018, 556, 191-203.
113. Ding, K.; Gulec, A.; Johnson, A. M.; et al. Identification of active sites in CO oxidation and water-gas shift over supported Pt catalysts. Science 2015, 350, 189-92.
114. Haruta, M. Gold catalysts prepared by coprecipitation for low-temperature oxidation of hydrogen and of carbon monoxide. J. Catal. 1989, 115, 301-9.
115. Green, I. X.; Tang, W.; Neurock, M.; Yates, J. T., Jr. Spectroscopic observation of dual catalytic sites during oxidation of CO on a Au/TiO2 catalyst. Science 2011, 333, 736-9.
116. Guzman, J.; Gates, B. C. Catalysis by supported gold: correlation between catalytic activity for CO oxidation and oxidation states of gold. J. Am. Chem. Soc. 2004, 126, 2672-3.
117. Zhao, K.; Qiao, B.; Wang, J.; Zhang, Y.; Zhang, T. A highly active and sintering-resistant Au/FeOx-hydroxyapatite catalyst for CO oxidation. Chem. Commun. 2011, 47, 1779-81.
118. Huang, J.; Wang, L.; Liu, Y.; Cao, Y.; He, H.; Fan, K. Gold nanoparticles supported on hydroxylapatite as high performance catalysts for low temperature CO oxidation. Appl. Catal. B. Environ. 2011, 101, 560-9.
119. Domínguez, M. I.; Romero-Sarria, F.; Centeno, M. A.; Odriozola, J. A. Gold/hydroxyapatite catalysts: synthesis, characterization and catalytic activity to CO oxidation. Appl. Catal. B. Environ. 2009, 87, 245-51.
120. Tang, H.; Wei, J.; Liu, F.; et al. Strong metal-support interactions between gold nanoparticles and nonoxides. J. Am. Chem. Soc. 2016, 138, 56-9.
121. Zhao, K.; Qiao, B.; Zhang, Y.; Wang, J. The roles of hydroxyapatite and FeOx in a Au/FeOx hydroxyapatite catalyst for CO oxidation. Chin. J. Catal. 2013, 34, 1386-94.
122. Guo, J.; Yu, H.; Dong, F.; Zhu, B.; Huang, W.; Zhang, S. High efficiency and stability of Au-Cu/hydroxyapatite catalyst for the oxidation of carbon monoxide. RSC. Adv. 2017, 7, 45420-31.
123. Boukha, Z.; Ayastuy, J. L.; Cortés-Reyes, M.; Alemany, L. J.; González-Velasco, J. R.; Gutiérrez-Ortiz, M. A. Catalytic performance of Cu/hydroxyapatite catalysts in CO preferential oxidation in H2-rich stream. Int. J. Hydrogen. Energy. 2019, 44, 12649-60.
125. Zhu, M.; Wachs, I. E. Iron-based catalysts for the high-temperature water-gas shift (HT-WGS) reaction: a review. ACS. Catal. 2016, 6, 722-32.
126. Miao, D.; Goldbach, A.; Xu, H. Platinum/apatite water-gas shift catalysts. ACS. Catal. 2016, 6, 775-83.
127. Ding, X.; Liu, W.; Zhao, J.; Wang, L.; Zou, Z. Photothermal CO2 catalysis toward the synthesis of solar fuel: from material and reactor engineering to techno-economic analysis. Adv. Mater. 2025, 37, e2312093.
128. Choi, S. H.; Song, I.; Dong, W. J. Recent progress of photothermal catalysts for carbon dioxide conversion. Energy. Mater. 2025, 5, 500062.
129. Medina, O. E.; Amell, A. A.; López, D.; Santamaría, A. Comprehensive review of nickel-based catalysts advancements for CO2 methanation. Renew. Sustain. Energy. Rev. 2025, 207, 114926.
130. Cheng, S.; Sun, Z.; Lim, K. H.; et al. Integrating plasmon and vacancies over oxide perovskite for synergistic CO2 methanation. Nano. Energy. 2025, 139, 110917.
131. Medeiros, F. G. M. D.; Ramalho, T. E. B.; Lotfi, S.; Rego de Vasconcelos, B. Direct flue gas hydrogenation to methane over hydroxyapatite-supported nickel catalyst. Fuel. Proc. Technol. 2023, 245, 107750.
132. Peng, Y.; Szalad, H.; Nikacevic, P.; et al. Co-doped hydroxyapatite as photothermal catalyst for selective CO2 hydrogenation. Appl. Catal. B. Environ. 2023, 333, 122790.
133. Guo, J.; Duchesne, P. N.; Wang, L.; et al. High-performance, scalable, and low-cost copper hydroxyapatite for photothermal CO2 reduction. ACS. Catal. 2020, 10, 13668-81.
134. Wai, M. H.; Ashok, J.; Dewangan, N.; et al. Influence of surface formate species on methane selectivity for carbon dioxide methanation over nickel hydroxyapatite catalyst. ChemCatChem 2020, 12, 6410-9.
135. Boukha, Z.; Bermejo-López, A.; Pereda-Ayo, B.; González-Marcos, J. A.; González-Velasco, J. R. Study on the promotional effect of lanthana addition on the performance of hydroxyapatite-supported Ni catalysts for the CO2 methanation reaction. Appl. Catal. B. Environ. 2022, 314, 121500.
136. Nguyen, T. T. V.; Phung Anh, N.; Ho, T. G.; et al. Hydroxyapatite derived from salmon bone as green ecoefficient support for ceria-doped nickel catalyst for CO2 methanation. ACS. Omega. 2022, 7, 36623-33.
137. Schiavoni, M.; Campisi, S.; Carniti, P.; Gervasini, A.; Delplanche, T. Focus on the catalytic performances of Cu-functionalized hydroxyapatites in NH3-SCR reaction. Appl. Catal. A. Gen. 2018, 563, 43-53.
138. Wang, J.; Liu, J.; Lu, Y.; Hong, D.; Yang, X. Catalytic performance of gold nanoparticles using different crystallinity HAP as carrier materials. Mater. Res. Bull. 2014, 55, 190-7.
139. Haruta, M.; Kobayashi, T.; Sano, H.; Yamada, N. Novel gold catalysts for the oxidation of carbon monoxide at a temperature far below 0 °C. Chem. Lett. 1987, 16, 405-8.
140. Haruta, M. Size- and support-dependency in the catalysis of gold. Catal. Today. 1997, 36, 153-66.
141. Kim, D.; Resasco, J.; Yu, Y.; Asiri, A. M.; Yang, P. Synergistic geometric and electronic effects for electrochemical reduction of carbon dioxide using gold-copper bimetallic nanoparticles. Nat. Commun. 2014, 5, 4948.
142. Boukha, Z.; Choya, A.; Cortés-Reyes, M.; et al. Influence of the calcination temperature on the activity of hydroxyapatite-supported palladium catalyst in the methane oxidation reaction. Appl. Catal. B. Environ. 2020, 277, 119280.
143. Kamieniak, J.; Kelly, P. J.; Doyle, A. M.; Banks, C. E. Influence of the metal/metal oxide redox cycle on the catalytic activity of methane oxidation over Pd and Ni doped hydroxyapatite. Catal. Commun. 2018, 107, 82-6.
144. Das, S.; Lim, K. H.; Gani, T. Z.; Aksari, S.; Kawi, S. Bi-functional CeO2 coated NiCo-MgAl core-shell catalyst with high activity and resistance to coke and H2S poisoning in methane dry reforming. Appl. Catal. B. Environ. 2023, 323, 122141.
145. Zhang, J.; Li, Y.; Song, H.; et al. Tuning metal-support interactions in nickel-zeolite catalysts leads to enhanced stability during dry reforming of methane. Nat. Commun. 2024, 15, 8566.
146. Kamieniak, J.; Kelly, P. J.; Banks, C. E.; Doyle, A. M. Methane emission management in a dual-fuel engine exhaust using Pd and Ni hydroxyapatite catalysts. Fuel 2017, 208, 314-20.
147. Campisi, S.; Galloni, M. G.; Marchetti, S. G.; Auroux, A.; Postole, G.; Gervasini, A. Functionalized iron hydroxyapatite as eco-friendly catalyst for NH3-SCR reaction: activity and role of iron speciation on the surface. ChemCatChem 2020, 12, 1676-90.
148. Mateo, D.; Cerrillo, J. L.; Durini, S.; Gascon, J. Fundamentals and applications of photo-thermal catalysis. Chem. Soc. Rev. 2021, 50, 2173-210.
149. Guo, J.; Liang, Y.; Song, R.; et al. Construction of new active sites: Cu substitution enabled surface frustrated lewis pairs over calcium hydroxyapatite for CO2 hydrogenation. Adv. Sci. 2021, 8, e2101382.
150. Li, B.; Chen, H.; Yuan, X. Influence of different La2O3 loading on hydroxyapatite supported nickel catalysts in the dry reforming of methane. Fuel 2024, 369, 131687.
151. Xiong, C.; Chen, S.; Yang, P.; Zha, S.; Zhao, Z.; Gong, J. Structure-performance relationships for propane dehydrogenation over aluminum supported vanadium oxide. ACS. Catal. 2019, 9, 5816-27.
152. Xin, Y.; Shirai, T. Noble-metal-free hydroxyapatite activated by facile mechanochemical treatment towards highly-efficient catalytic oxidation of volatile organic compound. Sci. Rep. 2021, 11, 7512.
153. Ye, R.; Ding, J.; Reina, T. R.; et al. Design of catalysts for selective CO2 hydrogenation. Nat. Synth. 2025, 4, 288-302.
154. Paris, C.; Karelovic, A.; Manrique, R.; et al. CO2 hydrogenation to methanol with Ga- and Zn-doped mesoporous Cu/SiO2 catalysts prepared by the aerosol-assisted sol-gel process. ChemSusChem 2020, 13, 6409-17.
155. Phan, T. S.; Pham Minh, D. New performing hydroxyapatite-based catalysts in dry-reforming of methane. Int. J. Hydrogen. Energy. 2023, 48, 30770-90.
156. Li, X.; Wang, Y.; Wei, X.; Zhao, Y. Effect of Na promoter on the catalytic performance of Pd-Cu/hydroxyapatite catalyst for room-temperature CO oxidation. Mol. Catal. 2020, 491, 111002.
157. Martínez-Hernández, H.; Mendoza-Nieto, J. A.; Pfeiffer, H.; Ortiz-Landeros, J.; Téllez-Jurado, L. Development of novel nano-hydroxyapatite doped with silver as effective catalysts for carbon monoxide oxidation. Chem. Eng. J. 2020, 401, 125992.
158. Maluf, S. S.; Tanabe, E. Y.; Nascente, P. A. P.; Assaf, E. M. Study of water - gas-shift reaction over La(1-y)SryNixCo(1-x)O3 perovskite as precursors. Top. Catal. 2011, 54, 210-8.
159. Rabee, A. I.; Cisneros, S.; Zhao, D.; et al. Uncovering the synergy between gold and sodium on ZrO2 for boosting the reverse water gas shift reaction: in-situ spectroscopic investigations. Appl. Catal. B. Environ. 2024, 345, 123685.
160. Rodriguez, J. A.; Grinter, D. C.; Liu, Z.; Palomino, R. M.; Senanayake, S. D. Ceria-based model catalysts: fundamental studies on the importance of the metal-ceria interface in CO oxidation, the water-gas shift, CO2 hydrogenation, and methane and alcohol reforming. Chem. Soc. Rev. 2017, 46, 1824-41.
161. Iriarte-Velasco, U.; Ayastuy, J. L.; Boukha, Z.; Bravo, R.; Gutierrez-Ortiz, MÁ. Transition metals supported on bone-derived hydroxyapatite as potential catalysts for the Water-Gas Shift reaction. Renew. Energy. 2018, 115, 641-8.
162. Su, T.; Gong, B.; Xie, X.; Luo, X.; Qin, Z.; Ji, H. Effect of cobalt on the activity of nickel-based/magnesium-substituted hydroxyapatite catalysts for dry reforming of methane. Chinese. J. Chem. Eng. 2024, 76, 281-91.
163. Boukha, Z.; Bermejo-López, A.; De-La-Torre, U.; González-Velasco, J. R. Behavior of nickel supported on calcium-enriched hydroxyapatite samples for CCU-methanation and ICCU-methanation processes. Appl. Catal. B. Environ. 2023, 338, 122989.
164. Pan, Q.; Peng, J.; Sun, T.; Wang, S.; Wang, S. Insight into the reaction route of CO2 methanation: promotion effect of medium basic sites. Catal. Commun. 2014, 45, 74-8.
165. Čapek, L.; Hájek, M.; Kutálek, P.; Smoláková, L. Aspects of potassium leaching in the heterogeneously catalyzed transesterification of rapeseed oil. Fuel 2014, 115, 443-51.
166. Nisar, J.; Razaq, R.; Farooq, M.; et al. Enhanced biodiesel production from Jatropha oil using calcined waste animal bones as catalyst. Renew. Energy. 2017, 101, 111-9.
167. Liu, C. A.; Cui, Y.; Zhou, Y. The recent progress of single-atom catalysts on amorphous substrates for electrocatalysis. Energy. Mater. 2025, 5, 500001.
168. Yamaguchi, K.; Mori, K.; Mizugaki, T.; Ebitani, K.; Kaneda, K. Creation of a monomeric Ru species on the surface of hydroxyapatite as an efficient heterogeneous catalyst for aerobic alcohol oxidation. J. Am. Chem. Soc. 2000, 122, 7144-5.
169. Ogo, S.; Maeda, S.; Sekine, Y. Coke resistance of Sr-hydroxyapatite supported Co catalyst for ethanol steam reforming. Chem. Lett. 2017, 46, 729-32.
170. Bittencourt, A. F. B.; Moraes, P. I. R.; Da Silva, J. L. F. Mechanistic insights into the direct partial oxidation of methane to methanol catalyzed by single-atom transition metals on hydroxyapatite. ACS. Omega. 2025, 10, 3868-77.
171. Tounsi, H.; Djemal, S.; Petitto, C.; Delahay, G. Copper loaded hydroxyapatite catalyst for selective catalytic reduction of nitric oxide with ammonia. Appl. Catal. B. Environ. 2011, 107, 158-63.
172. Lu, Z.; Cheng, Y.; Ma, D.; et al. Substitution of Fe in hydroxyapatite as an efficient single-atom catalyst for oxygen reduction reaction in biofuel cells: a first-principles study. Appl. Surf. Sci. 2021, 539, 148233.
173. Wang, J.; Li, W.; Di, A.; et al. Boosting C6-12 higher alcohols yield from ethanol upgrading over Ca-deficient mesoporous hydroxyapatite through intermediates enrichment of C4 aldehyde. ACS. Catal. 2025, 15, 8599-610.
174. Kamieniak, J.; Doyle, A. M.; Kelly, P. J.; Banks, C. E. Novel synthesis of mesoporous hydroxyapatite using carbon nanorods as a hard-template. Ceram. Int. 2017, 43, 5412-6.
175. Chingakham, C.; Tiwary, C.; Sajith, V. Waste animal bone as a novel layered heterogeneous catalyst for the transesterification of biodiesel. Catal. Lett. 2019, 149, 1100-10.
176. Bittencourt, A. F. B.; Valença, G. P.; Da Silva, J. L. F. Elucidating the catalytic valorization of ethanol over hydroxyapatite for sustainable butanol production: a first-principles mechanistic study. J. Phys. Chem. C. 2024, 128, 14663-73.
177. Wang, J.; Yan, X.; Wang, X.; Yang, M.; Xu, D. Selective activation of methane on hydroxyapatite surfaces: Insights from machine learning and density functional theory. Nano. Energy. 2024, 127, 109762.
178. Brasil, H.; Bittencourt, A. F.; Yokoo, K. C.; et al. Synthesis modification of hydroxyapatite surface for ethanol conversion: The role of the acidic/basic sites ratio. J. Catal. 2021, 404, 802-13.
179. Cheng, X.; He, Q.; Li, J.; Huang, Z.; Chi, R. Control of pore size of the bubble-template porous carbonated hydroxyapatite microsphere by adjustable pressure. Cryst. Growth. Des. 2009, 9, 2770-5.
180. Nowicki, D. A. Utilisation of carbon dioxide in the synthesis of multifunctional AB-type carbonated hydroxyapatite compositions. J. Solid. State. Chem. 2024, 334, 124678.
181. Ojeda-Niño, O. H.; Blanco, C.; Daza, C. E. High temperature CO2 capture of hydroxyapatite extracted from tilapia scales. Univ. Sci. 2017, 22, 215-36.
182. Chen, F. F.; Zhu, Y. J.; Xiong, Z. C.; Sun, T. W. Hydroxyapatite nanowires@metal-organic framework core/shell nanofibers: templated synthesis, peroxidase-like activity, and derived flexible recyclable test paper. Chemistry 2017, 23, 3328-37.
183. Sun, T. W.; Yu, W. L.; Zhu, Y. J.; et al. Hydroxyapatite nanowire@magnesium silicate core-shell hierarchical nanocomposite: synthesis and application in bone regeneration. ACS. Appl. Mater. Interfaces. 2017, 9, 16435-47.
184. Yang, Y. H.; Liu, C. H.; Liang, Y. H.; Lin, F. H.; Wu, K. C. Hollow mesoporous hydroxyapatite nanoparticles (hmHANPs) with enhanced drug loading and pH-responsive release properties for intracellular drug delivery. J. Mater. Chem. B. 2013, 1, 2447-50.
185. Omodolor, I. S.; Otor, H. O.; Andonegui, J. A.; Allen, B. J.; Alba-Rubio, A. C. Dual-function materials for CO2 capture and conversion: a review. Ind. Eng. Chem. Res. 2020, 59, 17612-31.
186. Annamalai, L.; Liu, Y.; Deshlahra, P. Selective C-H bond activation via NOx-mediated generation of strong H-abstractors. ACS. Catal. 2019, 9, 10324-38.
187. Liang, Y.; Li, Z.; Nourdine, M.; Shahid, S.; Takanabe, K. Methane coupling reaction in an oxy-steam stream through an OH radical pathway by using supported alkali metal catalysts. ChemCatChem 2014, 6, 1245-51.
188. Tu, S.; Guo, Y.; Zhang, Y.; et al. Piezocatalysis and piezo-photocatalysis: catalysts classification and modification strategy, reaction mechanism, and practical application. Adv. Funct. Mater. 2020, 30, 2005158.
189. Rodriguez, R.; Rangel, D.; Fonseca, G.; Gonzalez, M.; Vargas, S. Piezoelectric properties of synthetic hydroxyapatite-based organic-inorganic hydrated materials. Results. Phys. 2016, 6, 925-32.
190. Vasquez-Sancho, F.; Abdollahi, A.; Damjanovic, D.; Catalan, G. Flexoelectricity in bones. Adv. Mater. 2018, 30, 1705316.
191. Zhou, Y.; Wang, H.; Liu, X.; et al. Direct piezocatalytic conversion of methane into alcohols over hydroxyapatite. Nano. Energy. 2021, 79, 105449.
192. Xue, M.; Jin, Z.; Yang, B.; Xia, C.; Zhu, G. Boosting higher alcohols selectivity via regulating basicity of Ni/hydroxyapatite in ethanol upgrading. ACS. Catal. 2024, 14, 12654-63.
Cite This Article

How to Cite
Download Citation
Export Citation File:
Type of Import
Tips on Downloading Citation
Citation Manager File Format
Type of Import
Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.
Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.
About This Article
Copyright
Data & Comments
Data

Comments
Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at [email protected].