2.1. Carbon-supported monometallic Pd catalysts.
The study of the different properties of the active phase of the Pd-based catalysts and how they affect the catalytic activity is the starting point to achieve the development of catalysts with high performance and stability.
Among the properties of catalysts, particle size is a key parameter that can be modified by the preparation method. Li et al. conducted a study to investigate the catalytic activity and FA conversion of Pd/C catalysts with varying Pd particle sizes (ranging from 2.1 to 4.5 nm) [
36]. They prepared these catalysts with controlled particle sizes using a wet impregnation method, employing NaBH
4 for the reduction of the metal precursor and using sodium citrate as a stabilizing agent. The particle size variation was achieved by adjusting the reduction temperature and the sodium citrate to PdCl
2 ratio. Their X-ray photoelectron spectroscopy (XPS) analysis revealed that as the particle size increased, the proportion of oxidized Pd (Pd
2+ and Pd
4+) species decreased, while the proportion of Pd
0 increased. The catalytic activities of these Pd/C catalysts were evaluated in FA dehydrogenation in the liquid phase, and the smaller particle size catalyst (2.1 nm) exhibited the best performance with a turnover frequency (TOF) of 835 h
-1. The enhanced catalytic activity of smaller Pd NPs was attributed to their higher dispersion and a larger proportion of positively charged Pd species, leading to increased coulombic interaction with formate ions.
The composition of the surface of metal nanoparticles is commonly recognized to consist of distinct site types, including high-coordination terrace atoms and low-coordination atoms found at edges and corners. In the case of Pd NPs, a cuboctahedron-shaped particle model with a cubic close-packed structure is often employed to understand the role of the different sites in the catalytic reaction [
37]. The results of the values of the TOFs calculated for the different sites (low and high coordinated surface atoms) and also for the total of the surface atoms show that all the surface sites act as active sites and participate in the FA decomposition reaction. That was the case of the study reported by Navlani-García et al., who precisely controlled the size of the Pd nanoparticles by synthesizing PVP-capped colloidal NPs by the polyol method [
38,
39]. Pd NPs between 2.7 - 5.5 nm were obtained and the catalytic ability of the Pd/C catalysts in the FA decomposition in the liquid phase was evaluated by monitoring the H
2 production in a closed liquid-phase system during 3 h at 30 ºC. The results of the catalytic activity for the samples with different average diameters of the Pd NPS are depicted in
Figure 2(a), showing a volcano-like relationship between particle diameter and H
2 production. To further analyze the structure-activity relationship in H
2 production from formic acid, calculations were performed for the catalysts with different particle sizes assuming the particles as regular cuboctahedral in shape with a cubic close-packed structure and adopting the model of full-shell NPs. The TOF values calculated on the basis of the surface decreased in the order of Pd/C(3) (3.9 nm) > Pd/C(5) (5.5 nm) > Pd/C(2) (3.6 nm) > Pd/C(4) (4.2 nm) > Pd/C(1) (2.7 nm). This observation implies that not all Pd surface sites exhibit identical catalytic activity in FA dehydrogenation. As illustrated in
Figure 2(b), the TOF on low-coordinated atoms demonstrates a significant dependence on NPs size. In contrast, when focusing solely on the highly coordinated terrace atoms as catalytically active sites, the normalized TOF was found to be independent.
It should be noted that the catalysts of these two studies [
36,
39] were synthesized following different protocols and also different experimental conditions were used to assess the catalytic performance. So, even though smaller nanoparticles are usually preferred in catalysis, different conclusions could be drawn under certain conditions.
The incorporation of heteroatoms in the carbon supports has been another widely investigated aspect for the design of the catalysts used in the dehydrogenation of FA.
In particular, the incorporation of nitrogen functional groups in different types of carbon materials (carbon xerogels [
40], activated carbon [
41,
42], hierarchically porous carbon [
43] mesoporous carbon [
44,
45]), etc.) and by following different approaches addressed to either grafting or doping the carbon material, has attracted great interest in the last years [
46,
47,
48]. The resulting N-containing materials display improved catalytic performance in the dehydrogenation of FA compared to the N-free counterpart catalysts, which is usually attributed to electronic and acid-base properties, and also to the anchoring ability of N functional groups.
Our research group developed N-doped carbon-supported catalysts prepared from a lignocellulosic biomass residue (hemp residue) [
41]. In that study, the nitrogen incorporation was carried out through an organic reaction and the surface nitrogen content was close to 2 at.%. Not only the role of the nitrogen functional groups was analyzed, but also the effect of the preparation of the catalysts was considered in that study. The results of the catalytic activity registered at 75 ºC indicated that the developed materials displayed excellent catalytic performance, with great stability at least during 6 consecutive reaction cycles and TOF of 8365 h
-1, expressed per surface Pd mol.
Yao et al. prepared Pd-based catalysts supported on N-doped activated carbon (denoted as HTNC) after a high-temperature amination process [
49]. The amination process was carried out by heating the activated carbon at different temperatures (from 650 to 1050 °C) in a NH
3 flow. XPS spectra of N 1s show different functional groups for all the catalysts (i.e. pyridinic N, pyrrole N, and graphitic N). The highest content of total nitrogen was observed in Pd/HTNC-950. The Pd 3d XPS spectra showed the presence of Pd
0 and Pd
2+ and it was seen that the content in Pd
2+ increased with the temperature used in the amination, which confirmed the interaction between Pd and nitrogen-doped activated carbon (AC). The catalytic performance of the as-synthesized catalysts was tested at 30 ºC in a FA/sodium formate (SF) 1:1 solution. The catalysts had remarkably different activities, and the best activity was obtained by Pd/HTNC-950 with a TOF of 1631 h
-1. These results confirmed that the presence of N species in the ACs was essential to enhance the catalytic activity in FA dehydrogenation. It was claimed that the beneficial effect of N groups was due to both the electronic properties of Pd species and the high dispersion of Pd NPs achieved in the catalysts. On the one hand, it was observed that N species could modify the electron state of Pd and stabilize Pd
2+ species, which interact with formate ions, key intermediates in FA dehydrogenation. Conversely, the HTNC effectively immobilized free Pd ions, which led to the creation of highly dispersed Pd nanoparticles. This process increased the prevalence of lower coordination sites in smaller particles, a factor that can be associated with improved performance in the reaction. [
50]. The stability of the catalysts was measured by 5 reaction cycles. Even after the fifth run, the catalytic activity of the Pd/HTNC-950 catalyst remained consistently high, underscoring the robustness of this catalyst.
Jeon et al. developed a method to dope the AC with N groups that involved the pyrolysis of dicyandiamide using commercial Ketjen black [
51]. The catalysts were prepared by a wet impregnation method with reduction assisted by NaBH
4. The results obtained by TEM showed that the Pd NPs were well dispersed on the N-C support with an average size of approximately 1-2 nm. The N 1s XPS spectra of Pd/N-C displayed the presence of C-N bonds (C=N+-C and C=N), attributed to the incorporation of nitrogen atoms into the carbon structure The Pd 3d XPS demonstrated that the Pd electron density of Pd/N-C slightly increased by the electrons transferred from N to Pd species through the interaction between the metal and support. The study of the catalytic activity of these catalysts was performed at 45 ºC and Pd/N-C exhibited a better performance than Pd/C obtaining a full conversion in 100 min. Due to the excellent catalytic activity described above, it was developed a fuel cell powering system connected to the FA-based H
2 generation reactor using Pd/N-C catalysts. For performing a 200 W PEMFC stack, 9.5 mL min
-1 of FA was continuously supplied to the catalytic reactor. The stack received a maximum current load of 8 A, accompanied by a lower cell voltage of around 25 V. This configuration yielded a power generation of 200 W through the utilization of fuel derived from FA dehydrogenation. To establish the durability and applicability of the integrated H
2 fuel cell stacks and the FA-based H
2 generator system employing Pd/N-C, long-term chronopotentiometric measurements were implemented. A hydrogen production rate of approximately 3 L min
-1 was achieved in galvanostatic conditions at a current of 6.45 A. The fuel cell demonstrated consistent performance for a duration of 80 minutes, with no discernible decline in stack voltage. This suggested the absence of a loss in catalytic activity for hydrogen production and indicated that there was no poisoning of the fuel cell anode catalysts, potentially caused by the formation of CO during the dehydrogenation of FA.
Chet et al. reported an efficient procedure to obtain ultrafine Pd NPs supported on activated N-doped carbon [
52]. The preparation procedure of nitrogen-doped activated carbons (NHPC-AC) was based on the mixing of activated carbon, NaHCO
3 and NH
4HCO
3 (mass ratio 1:3.3) and then calcination at 700 ºC for 1 h in N
2 flow. TEM images for the Pd/NHPC-AC catalyst indicated a good dispersion of the NPs and an average size of 1.88 nm, being smaller than for Pd/HPC-AC (2.56 nm). XPS analysis revealed the presence of pyridinic, pyrrolic, and graphitic nitrogen. The catalytic activity measured at 60 ºC with a FA:SF molar ratio of 1/1 demonstrated better behavior for Pd/HPC-NAC than for Pd/HPC-AC, reaching TOF values of 4115 h
-1 at 60 ºC and 1910 h
-1 at 30 ºC. The influence of SF was examined by varying the molar ratio of FA/SF. As the FA/SF molar ratio decreased, there was a significant enhancement in catalytic activity, reaching a plateau when the molar ratio fell below 1:1. Moreover, even after undergoing recycling for five cycles, Pd/NHPC-AC displayed sustained high catalytic activity, and the average size of the Pd nanoparticles NPs remained small.
Shao et al. recently investigated a method to prepare amine-functionalized hierarchically porous carbon (TPC-NH
2) as a support for Pd NPs which obtained excellent results. The desired architecture of the catalysts was achieved by choosing MgO as a template and sodium hypophosphite to widen the pores achieving the hierarchically porous carbon (TPC). Then, the TPCs were functionalized with amino groups and the catalysts were prepared by a wet impregnation method with NaBH
4 reduction. With this method, ultrafine Pd NPs with an average size of 2.1 nm were obtained for Pd/NH
2-TPC. EDS elemental mapping analysis showed a uniform distribution of the elements C, N, O, Pd and Si. The catalytic performance of Pd/NH
2-TPC at 25ºC using a 1 M FA solution without additives, revealed a high activity reaching a TOF value of 4312 h
-1 and the increase of temperature to 60 ºC achieved a TOF of 12864 h
-1 (see
Figure 3). These results, compared with the rest of the catalysts reported for this application, using a pure FA solution without additives, show the best TOF values. The excellent catalytic activity results obtained were related to an increase of the reaction rate due to the enhancement of the mass transfer promoted by the hierarchical porous structure.
The positive effect of the presence of nitrogen functional groups has not only been observed when the reaction is performed in the liquid phase, but interesting results have also been achieved for the decomposition of FA in the gas phase. That was the case of Bulushev et al. who reported a study where the preparation of single Pd
2+ cations supported on N-doped carbon (Pd/N-CM) was achieved [
54]. The presence of these single sites was confirmed through aberration-corrected scanning transmission electron microscopy (ac STEM). The results of the catalytic activity measured in the gas phase for Pd/N-CM at 100 ºC showed a TOF value of 0.121 s
-1. The results of X-ray photoelectron spectroscopy (XPS) and near-edge X-ray absorption fine structure (NEXAFS) studies after ex-situ reduction in hydrogen at 300 ºC, showed that the Pd species exist in a Pd
2+ state coordinated by nitrogen species within the support. Additionally, extended density functional theory (DFT) calculations validated the critical role of an isolated Pd atom as the active site for FA decomposition, resulting in the production of an adsorbed hydrogen atom and a carboxyl fragment. However, this activity was only viable when the Pd atoms were coordinated by a pair of pyridinic-type nitrogen atoms located at the open edge of the graphene sheet. Thus, the nitrogen doping of the carbon support plays a pivotal role in creating and stabilizing these new active Pd sites. Bulushev also reported the synthesis of catalysts based on Pd active sites on covalent triazine frameworks [
55]. Pd catalysts were prepared by impregnation with the supports: 5,6,11,12,17,18-hexaazatrinaphthylene-2,8,14-tricarbonitrile (hatnCTF), 4,4′-malonyldibenzonitrile (acacCTF) and g-C
3N
4. The surface N/C ratio measured by XPS was much lower in Pd/acacCTF compared to Pd/hatnCTF. Similarly, the surface oxygen content is 3 times higher in Pd/acacCTF than for Pd/hatnCTF. The surface concentration ratio of pyridinic nitrogen to the overall nitrogen surface concentration was approximately 40% in both samples. However, in the Pd/g-C
3N
4 sample, nitrogen is predominantly found in the pyridinic configuration. By means of HRTEM and HAADF/STEM the formation of nanoparticles was observed in Pd/acacCTF and Pd/g-C
3N
4 with an average size of 5.2 and 3.0 nm, respectively, while for Pd/hatnCTF only single atoms were observed and for Pd/acCTF the presence of these could also be appreciated. Nevertheless, Pd/g-C
3N
4 only presented Pd in the form of NPs. The Pd/hatnCTF sample presented a higher proportion of Pd
2+ sites than the rest of the samples, which was confirmed by both XPS and XANES. After treatment of the spectra obtained by EXAFS it was shown that the single atoms in Pd/hatnCTF were found as Pd
2+-C
1N
3 sites and the Pd/acacCTF sample contained single-atom Pd
2+-O
4 sites. The presence of a significant proportion of metal NPs in Pd/acacCTF was also confirmed. The catalytic activity measured in gas phase showed the best performance for Pd/hatnCTF with a TOF of 0.28 s
-1 at 180 °C. Additionally, a stability assessment was conducted for Pd/hatnCTF that showed no deactivation at 180 °C, and there was even a slight increase in FA conversion within the initial 5 hours of the test. Those results could be attributed to the presence of single-atom Pd sites (Pd-C
1N
3) in Pd/hatnCTF, while in Pd/acCTF the Pd single-atom Pd-O
4 sites are not active for the reaction; thus, the low activity observed for this catalyst is only due to the presence of NPs.
2.2. Carbon-supported multimetallic Pd-based catalysts.
Exceptional outcomes have been attained with Pd-based bimetallic catalysts designed as alloys or core-shell structures. In these configurations, the electronic promotion of Pd species occurs through charge transfer from the other component. The favorable effects of the construction of multimetallic catalysts extend to mitigating deactivation caused by poisonous intermediates such as CO. Additionally, these catalysts weaken hydrogen adsorption on the surface of the NPs, thereby promoting the favored combination of two hydrogen atoms to form an H
2 molecule [
56].
Specifically, extensive research has been conducted on systems incorporating Pd in conjunction with other noble metals. Pd-Ag-based catalysts, in particular, have remarkable performance. This can be attributed to the electron-rich Pd species resulting from electron transfer from Ag to Pd by the difference in electronegativity (2.2 and 1.9 in the Pauling scale, for Pd and Ag, respectively) [
57].
Navlani-García et al. studied colloidal bimetallic PdAg NPs that were prepared by using the polyol method and supported in a commercial activated carbon [
58]. The study was performed with different Pd/Ag ratios, as well as different PVP/metal ratios. TEM micrographs showed how the addition of Ag in small proportions (Pd1Ag0.5, Pd1Ag1, Pd1Ag2) reduced the particle size compared to monometallic catalysts. The catalytic performance of PdAg/C catalysts in the FA dehydrogenation reaction was evaluated by monitoring the H
2 produced during 3 h at 30 ºC. The high influence of the Pd/Ag ratio as well as the PVP/metal ratio could be appreciated, being the catalyst with the best activity Pd1Ag2 with PVP/M=1. EXAFS measurements for Pd1Ag2/C confirmed the heteroatomic Pd-Ag bonding in the studied catalysts. The analysis of XPS spectra indicated that Pd and Ag were predominantly present in metallic form, but a minor presence of oxidized forms (Pd
2+ and Ag
+) was observed for both elements. The presence of oxidized species was assigned to the electron-withdrawing property of PVP through the C=O groups, resulting in electron-deficient metal surfaces.
Recently, Kim et al. investigated different Pd:Ag compositions for elucidating the alloying effect on PdAg NPs supported on carbon nanotubes (CNT) catalysts [
59]. The catalysts were prepared by a deposition-precipitation method using CNT as support. The total content of Pd and Ag was fixed at 5 wt.% while varying their molar ratios. The Pd binding energy obtained by XPS showed the highest shift to lower energy for the catalyst with an Ag/Pd ratio of 3/7, which can be attributed to increased interaction between the metals. The initial TOF values showed a volcano-type curve according to the Ag/Pd ratio, the Pd
7Ag
3/CNT catalyst presented the highest value. In this sense, kinetic isotopic effect (KIE) experiments were carried out to investigate the alloy effect. The KIE results showed lower values with increasing Ag content, reaching a minimum where it remains constant for Pd/Ag molar ratios of 7:3. This lower KIE (k
h/k
d) value can be associated with a weaker C-H bond, which facilitates the breaking of this bond of the adsorbed HCOO
-, thereby easing the course overall reaction. In that study, the alloying effect of Ag on Pd was explained as an electronic modification of Pd, which excited the molecular orbital of the adsorbed HCOO
− to have a weaker C–H bond. Consequently, this facilitated the cleavage of the C–H bond in HCOO
−*, leading to an increased activity in the dehydrogenation of FA over the PdAg/CNT catalysts.
The positive effect of the incorporation of nitrogen functional groups has also been evidenced for bimetallic PdAg catalysts. For instance, Nabid et al. reported on the utilization of Ag-core Pd-shell NPs supported on nitrogen-doped graphene carbon nanotube aerogel (Ag@Pd/N-GCNT) [
60]. In that investigation, graphene and CNTs were incorporated into a N-doped hybrid aerogel. That aerogel was then sequentially loaded with Ag and Pd in order to achieve controlled composition of the core-shell NPs. The findings revealed that, beyond the nitrogen sites, the graphene carbon nanotube aerogel provided additional stabilization points for the NPs, preventing their aggregation. Catalysts with an Ag/Pd molar ratio of 1/1 showcased the highest activity among the examined compositions, achieving a TOF value of 413 h
-1 at 25 ºC.
In our research group, catalysts based on Pd and PdAg NPs were designed from activated carbons prepared from a biomass residue (almond shell) and doped with nitrogen groups [
61]. Two sets of catalysts were prepared, some prepared by a wet impregnation method with reduction by NaBH
4 and others in which the reduction step with NaBH
4 was omitted and the metals were expected to be reduced in-situ during the reaction. The results of the catalytic activity measured at 75 °C showed significant differences between the pre-reduced catalysts and those that were not. The reduced PdAg catalysts showed better activity than the monometallic counterparts. In the non-pre-reduced bimetallic catalysts, an induction period was observed at the beginning of the gas-time curve. PdAg/N-AS showed the best performance achieving an initial TOF of 1577 h
−1 and the good stability of the catalyst was demonstrated after 6 consecutive reaction cycles.
Furthermore, for Pd bimetallic catalysts, other noble metals such as Au have been used to form PdAu alloys for the dehydrogenation of FA. Jiang et al. reported a study in which AuPd NPs supported on N-doped carbon nanosheets (n-CNS) were assessed [
62]. The synthesis procedure is described in
Figure 4 and it consists first of a hydrothermal treatment of a solution of graphitic carbon nitride (g-C
3N
4) and glucose followed by calcination, using different temperatures. AuPd catalysts were prepared by wet impregnation with a reduction step. The images obtained by HAADF-STEM and STEM-EDX confirmed the homogeneous distribution of N and it was also found that the signal for Au an Pd are at the same position, which strongly confirms de existence of AuPd alloy. The highest activity was obtained for AuPd/n-CNS-Th-160 (hydrothermal treatment at 160 ºC) which achieved a TOF of 459 h
-1 at 25 ºC and a record value of 1896 h
-1 at 60 ºC with 100% selectivity to CO
2 and H
2. Furthermore, the study demonstrated that the catalytic activity did not exhibit a direct correlation with the mass of nitrogen dopant. Instead, the nitrogen bonding configurations played a crucial role, and a higher ratio of graphitic N to pyridinic N was shown to modify the electron structure distribution of the metal and enhance the interaction between AuPd metal nanoparticles and the support.
Hong et al. developed a simple method for the immobilization of PdAu NPs on commercial carbons that resulted in excellent catalytic results [
63]. The method consisted of wet mixing of the metal precursors, using L-arginine (LA) as mediating reagent, with MSC-30 and Vulcan XC-72R and a reduction step. The catalyst formed by PdAu NPs supported on MSC-30 was denoted as Pd
1Au
1/30-LA and the one supported on Vulcan XC-72R was denoted as Pd
1Au
1/72-LA. Severe differences in the catalysts were observed. Firstly, the mean particle size obtained by TEM was smaller for Pd
1Au
1/72-LA suggesting the importance of the metal-support interaction. Pd
1Au
1/30-LA showed exceptional performance for the dehydrogenation of FA using a pure FA solution at 60 ºC reaching a TOF value of 8355 h
-1. Moreover, the stability of that catalyst was studied after 5 consecutive reaction cycles which revealed its excellent durability. However, when the reaction was studied using a 1:3 molar FA/SF solution the catalyst with the best activity was Pd
1Au
1/72-LA obtaining a TOF of 11958 h
-1 at 60 ºC, and also the stability test proved its high durability. In addition, LA was found to play an important role in modulating the size and dispersion of NPs.
Noble metal-based architectures are usually preferred for their reported catalytic properties. However, the scarcity and high cost of these metals have motivated the search for alternative non-noble metals like transition metals such as Co, Ni, Cu, etc.
Navlani-García et al. studied PdCo-based bimetallic systems with different Pd/Co ratios supported on (
g-C
3N
4) [
64]. The Pd-based catalysts were prepared by the standard impregnation method, mixing
g-C
3N
4 with the appropriate amount of Na
2PdCl
4 and Co(NO
3)
2 to prepare catalysts with several Pd/Co molar ratios (Pd/Co = 1/0.1, 1/0.25, 1/0.7, 1/1 and 1.15). The catalytic activity evaluated at 75 ºC for the different compositions of the NPs showed that the gradual addition of Co resulted in outperforming catalysts until the optimal composition was reached in the sample PdCo/
g-C
3N
4 (1/0.7) achieving a TOF value of 1193 h
-1. The additional introduction of Co was proved to be ineffective and counterproductive in terms of FA dehydrogenation capability, that could be attributed to the partial shielding of the active sites in the Pd NPs by the excess of cobalt. The XPS spectra of Pd 3d revealed a slight shift toward lower binding energies in the PdCo catalysts compared to monometallic Pd. That shift confirmed the existence of alloyed NPs and suggested the formation of electron-rich Pd species on the PdCo particles, which could attributed to the donation of electrons from Co to Pd, as indicated by their respective electronegativities (values in Pauling units of 1.88 and 2.20, respectively). The positive influence of the alloy system in catalyzing FA dehydrogenation was substantiated through potential energy profiles generated by DFT calculations, utilizing Pd
43 and Pd
22Co
21 clusters as models for monometallic and alloy NPs, respectively. Considering the suggested reaction mechanism (
Figure 5), the lower energy barriers observed for the Pd
22Co
21 cluster, in comparison to the monometallic Pd43, confirmed the favorable impact of bimetallic NPs in enhancing the FA dehydrogenation reaction.
Tamarany et al. studied systems based on PdNi supported in N-doped carbons [
65]. The N-doped carbons (N-C) were obtained by mixing dicyanoamide with carbon black at 100 °C and then pyrolyzed at 550 °C for 4 h in N
2 flow. Pd
1Ni
x alloys supported on N–C (Pd
1Ni
x/N–C) were prepared by dispersing N-C into aqueous solutions containing Pd(NO
3)
2·2H
2O and Ni(NO
3)
2·6H
2O at different Pd/Ni mole ratios (1/0.33, 1/1 and 1/3). The resulting black powders were collected and reduced under a H
2/N
2 flow (20% H
2) at 450 ºC for 4 h, affording Pd
1Ni
x/N–C. The catalytic activity for FA dehydrogenation was assessed at 30 °C with FA:SF solution in 1/1 molar ratio for 3 catalysts with different Pd/Ni ratios. The results showed a volcano-type behavior where the maximum catalytic activity for Pd
1Ni
1.3/N–C produced a TOF value of 447 h
-1. In another experiment, the influence of temperature on the dehydrogenation activity with Pd
1Ni
1.3/N–C was studied, obtaining an initial TOF value of 4804 h
-1 at 65 ºC. That value could be attributed in part to the fact that the sodium formate can potentially react with water to produce hydrogen (HCOONa + H
2O → NaHCO
3 + H
2), particularly at temperatures higher than 65 ºC. Furthermore, the formation of PdNi alloys was confirmed by: a) elemental mapping of Pd
1Ni
1.3/N–C by HAADF-STEM through the overlap in the positions of Pd and Ni, and b) by the shift to lower binding energies of Pd 3d spectra obtained by XPS.
Mori et al. investigated PdCu alloy NPs supported on a macroreticular basic resin with -N(CH
3)
2 groups [
66]. The deposition of both Pd and Cu within the basic resin was performed using a simple ion exchange method from an aqueous solution of PdCl
2 and CuNO
3·3H
2O, and later the samples were reduced by NaBH
4. EXAFS analysis revealed the presence of heteroatomic Pd-Cu bonding. The formation of PdCu nanoparticles with an average size of 1.9 nm was confirmed by TEM. The catalytic activity results measured for different Pd/Cu molar ratios show a volcano-type trend, where the Pd50Cu50 composition presents the maximum TOF value, 810 h
-1 at 75 ºC using a FA/SF 9:1 aqueous solution. The improvement in catalytic activity of the bimetallic PdCu catalyst over the monometallic Pd catalyst can be explained by the energetic efficiency of transferring Cu electrons to Pd atoms due to the difference in ionization potential (Cu: 7.72 eV and Pd 8.34 eV). Additionally, that study confirms the significant impact of the basicity of the resin on the breaking of O-H bonds and the introduction of Cu is shown to influence the rate-determining step involving C-H bond dissociation from the metal-formate intermediate, attributable to the formation of electron-rich Pd species.
Pd-based trimetallic catalysts have not been as extensively studied as bimetallic catalysts, but they are an interesting alternative to achieve a Pd content reduction and modulate the electronic properties of Pd species.
Yurderi et al. designed carbon-supported trimetallic PdNiAg NPs as catalysts for FA dehydrogenation [
67]. The preparation of the catalysts was achieved by a wet impregnation method with a reduction step and the precursors of the metals were mixed in different molar ratios with the commercial activated carbon. The morphology and size of the PdNiAg/C NPs were studied by TEM, obtaining a good dispersion and with an average particle size of 5.6 nm. XPS spectra showed the presence of metallic NPs and also oxidized species such as PdO, NiO, Ni(OH)
2, NiOOH and Ag
2O. Among the different compositions studied, the Pd
0.58Ni
0.18Ag
0.24/C catalyst shows the best catalytic activity, reaching a TOF of 85 h
-1 at 50 ºC in a 1:1 FA/SF solution. The reusability of that catalyst was studied after five consecutive cycles of reaction and it was demonstrated that 94% of the initial catalytic activity was retained.
Wang et al. prepared NiAuPd alloy supported on a commercial carbon [
68]. The synthesis of the catalysts was carried out by a co-reduction method without any surfactant at 25 °C. The images obtained by HAADF-STEM showed that Ni, Au, and Pd were homogeneously distributed in each particle, pointing out that the alloy structure is indeed formed. The catalytic activity results for Ni
0.40Au
0.15Pd
0.45/C revealed a TOF of 12.4 h
-1 (calculated on the basis of the total amount of metal) at 25 °C and in a pure FA solution without additives.
Dong et al. tailored a system based on PdCoNi NPs supported on N-CN. N-CN was prepared by a combination of CN and 3-aminopropyl triethoxysilane [
69]. The catalysts based on the PdCoNi NPs were synthesized by a co-reduction method. Ultrafine NPs were obtained for the Pd
0.6 Co
0.2Ni
0.2/N-CN catalyst with an average size of 1.6 nm. XPS analysis detected the electron transfer from Co and Ni to Pd, indicating the formation of a solid solution structure of Pd
0.6Co
0.2Ni
0.2 NPs, and the presence of pyrrolic N was also detected. The catalytic activity of Pd
0.6Co
0.2Ni
0.2/N-CN was measured at 25°C in a 1:1 aqueous FA/SF solution obtaining an initial TOF value of 1249 h
-1.
Liu et al. performed a study based on trimetallic AuPdIr nanoalloy supported on N-doped reduced graphene oxide (N-GO) [
70]. AuPdIr catalysts were prepared by a co-reduction method by mixing the metal precursors with an APTS/GO solution, obtaining Au
0.35Pd
0.5Ir
0.15/NH
2-N-rGO. By XPS, the electron transfer from Pd and Ir to Au has been observed, which is consistent with their electronegativity (Au 2.4, Pd 2.2, Ir 2.2), thus confirming the nanoalloy formation. The performance in FA dehydrogenation was tested obtaining a TOF value of 12781.2 h
-1 at 25 °C in a pure FA solution (
Figure 6A), and the temperature increase to 60 °C gave a TOF value of 36598.4 h
-1. The recycling stability of Au
0.35Pd
0.5Ir
0.15/NH
2-N-rGO was tested by adding an extra aliquot of FA dilution after completion of the previous one and showed high stability after 10 reaction cycles. To further test the long-term stability, a 150-day experiment was designed. As depicted in
Figure 6D, the catalyst retained the ability to catalyze 100% FA dehydrogenation for 7.08 minutes at room temperature, even after 150 days. According to DFT calculations, that robust activity was ascribed to the incorporation of the high surface energy element iridium (Ir), which alters the initial adsorption configuration of HCOOH* and improves the overall performance of the reaction.