Identification of Ni3Fe alloy as a candidate catalyst for quinoline selective hydrogenation with computations†
Received
8th September 2024
, Accepted 21st October 2024
First published on 22nd October 2024
Abstract
The 1,2,3,4-tetrahydroquinoline (py-THQL) is a crucial intermediate and fragment in chemical synthesis, but its production from quinoline (QL) selective hydrogenation in heterogeneous catalysis mainly relies on noble-metal-based catalysts. Therefore, the design of catalysts composed of earth-abundant elements for this reaction is meaningful. In this work, using density functional theory (DFT) calculations, we found the binding energy of QL to be a suitable descriptor to illustrate the general activity trend of metallic catalysts for QL hydrogenation. Among the screened bimetallic alloys composed of Fe, Co, Ni, and Cu, we computationally identified Ni3Fe as a promising candidate catalyst with high stability, while our systematic mechanistic calculations showed the low energy barriers for each hydrogenation step. Our established DFT-based mean-field microkinetic model indicates a much higher turnover frequency for py-THQL production on the Ni3Fe(111) surface than on the experimentally reported high-performance AuPd3(111) surface. This work not only identified a valuable descriptor for the rational catalyst screening for the complex QL hydrogenation reaction but also theoretically predicted a cost-effective Ni3Fe catalyst for the hydrogenation reaction.
1. Introduction
Nitrogen-containing compounds play pivotal roles in the pharmaceutical and agrochemical industries. Among them, 1,2,3,4-tetrahydroquinoline (py-THQL) and 5,6,7,8-tetrahydroquinoline (bz-THQL) are widely used as key intermediates and fragments in chemical synthesis.1 In particular, the asymmetric structure and saturated N-heterocycle of py-THQL in various alkaloids and medicines enable its high chemical and biological activities against pathogenic bacteria and cancer cells.2–4 Meanwhile, the py-THQL is also a crucial structure unit in dyes, working as an efficient sensitizer for dyes-sensitized solar cells (DSSCs).5–7 Consequently, developing practical approaches for the efficient synthesis of py-THQL is meaningful for the aforementioned fields. Currently, there are three primary methods for py-THQL synthesis, including catalytic cyclization,8 Beckman rearrangement,9 and quinoline (QL) selective semi-hydrogenation.4,10,11 Besides, the synthesis of py-THQL and derivatives through cross-coupling of quinoline with organic molecules could be achieved by using the electrochemical reductive method.12,13 Among them, the selective semi-hydrogenation of QL using H2 as the hydrogen source has the advantages of being a green process and having high atomic efficiency.14 However, the selectivity control to avoid the formation of bz-THQL and decahydroquinoline (DHQL), as well as the harsh reaction conditions in this process, are the key challenges to be tackled,15–17 which call for the design of high-performance catalysts.
The research on QL selective hydrogenation can be traced back to 1982, Fish et al.10 identified Ru-based homogeneous catalysts to be active for this process, followed by the discoveries of Au,18 Rh,19 and Ir20 complexes. Meanwhile, Pd, Pt, Ru, Co and Rh-based heterogeneous catalysts were also reported to exhibit outstanding performance in the QL hydrogenation reaction.11,21–26 Apart from pure metals, alloying has been recently found to be an effective strategy to promote this reaction, and a group of alloys was reported to be promising, such as Au–Pd,14 Pd–Ni,27 Cu–Ru,28 Ni–Ir,29 Ni–Co,30 Au5Pt,31 and AuPd3.32 Compared with the wide investigations experimentally, theoretical studies on the QL hydrogenation reaction have been relatively scarce due to the complex reaction mechanism. Xu and coworkers33 studied QL hydrogenation to py-THQL on Ni-promoted MoS2 with density functional theory (DFT) calculations, where the reaction pathway followed the sequential hydrogenation of the N4, C3, C2, and C1 atoms on the N-heterocycle (atomic labels shown in Scheme 1). Similarly, Wang and coworkers34 studied the QL hydrogenation mechanism on MoP(111) surface and found that the formation of py-THQL occurs via sequential hydrogenation of C3, C2, C1, and N4 atoms.
 |
| | Scheme 1 Natural products, medicines, alkaloids, and dyes containing py-THQL framework, as well as the possible hydrogenation sites labeled as C1, C2, C3, N4, C1′, C2′, C3′, C4′. | |
Indeed, the high complexity of the QL hydrogenation process (Fig. S1†) makes it computationally very expensive to conduct mechanism calculations on large numbers of catalysts and eventually hampers the rational design of earth-abundant catalysts for this reaction with computations. In this respect, the Sabatier principle capable of guiding the rational catalyst design for many reactions35–41 will make efficient catalyst screening possible if adequate descriptors for QL hydrogenation can be identified.
In this work, we theoretically identified the binding energy of QL (EQL) on metallic catalysts as an adequate descriptor for determining the activity of QL hydrogenation, where a volcano plot was established based on our previous joint experimental and theoretical studies of QL hydrogenation on AuPd-based alloys.32 Then, we calculated the binding energy of QL on 18 bimetallic alloys composed of earth-abundant Fe, Co, Ni, and Cu elements. Among them, the Ni3Fe alloy was theoretically predicted to be promising for QL hydrogenation. Our systematic DFT calculations and mean field microkinetic modeling (MKM) analysis showed that this Ni3Fe catalyst exhibited higher TOF for py-THQL production than the AuPd3 catalyst under the same reaction conditions. This work provides the theoretical feasibility of using NiFe-based alloy for QL selective hydrogenation.
2. Computational methods
All DFT calculations are implemented in the Vienna ab initio simulation package (VASP).42 The BEEF-vdW functional43 was used to describe the exchange-correlation contribution to the electronic energy due to its relatively high accuracy in describing the adsorption proprieties of adsorbates on transition metal surfaces. We employed the periodic surface slab model to simulate the adsorption and the hydrogenation process. The most stable (111) surface for FCC and (110) surface for BCC were chosen for the mechanistic study. The surface model is a 4-layered (4 × 4) supercell constrained bottom two layers and relaxed top two layers with a vacuum thickness of 15 Å to eliminate the interactions among the periodic surfaces. The cutoff energy was set to 400 eV to obtain accurate energy. All calculations were done when the force and energy difference were smaller than 0.05 eV Å−1 and 10−4 eV, respectively. The binding energy of QL is defined as EQL = E(QL/surface) − E(surface) − E(QL), where E(QL/surface), E(surface), and E(QL) are the total electronic energies of QL adsorbed on bimetallic catalysts surfaces, clean surfaces, and QL molecule in the gas phase, respectively. The transition states of each hydrogenation step are located by using the climbing image nudged elastic band (CINEB) method.44 The Gibbs free energy barrier (Ga) is defined as the Ga = G(TS) − G(IS), where G(IS), G(TS) are the free energies of initial states (IS) and transition state (TS) of each hydrogenation steps, respectively. All transition states are analyzed by frequency calculations with only one imaginary frequency. The thermodynamic corrections of adsorbates and gases are based on harmonic and ideal gas approximations as implemented in ASE.45
3. Results and discussion
3.1. Sabatier principle for QL hydrogenation
To achieve an efficient screening of catalysts for QL hydrogenation, it is necessary to establish a rational activity trend for a group of well-studied catalysts, ideally from experimental results. Fortunately, we had the experimentally measured values of turnover frequencies (TOF) for QL hydrogenation on Au, AuPd3, AuPd, Au3Pd, and Pd in our previous joint experimental and theoretical studies.32 Meanwhile, our mechanistic studies showed that the binding energy of QL on the metallic catalyst could work as a suitable descriptor to describe the general activity trend of the aforementioned AuPd-based catalysts. On this basis, the pure Pd catalyst with a too-strong binding of QL shows limited activity due to the surface poising, while the Au with a too-weak binding has difficulty activating reactants and eventually shows low activity. The AuPd-based alloys with moderate binding strength of QL show improved activity compared with the pure Au and Pd metals. Indeed, this follows the Sabatier principle well, which provides valuable guidance for our efficient screening of candidate catalysts for QL hydrogenation. As shown in Fig. S2,† we theoretically fitted the Sabatier volcano between EQL and TOF of QL hydrogenation. Theoretically, the catalyst with an adequate EQL may perform well for QL hydrogenation.
To design cost-effective catalysts for QL hydrogenation reaction, we focused on the alloys composed of earth-abundant Fe, Co, Ni, and Cu elements at three compositions (1
:
1, 1
:
2, and 1
:
3). Then, the close-packed surface of these 18 alloys were chosen as models to calculate the binding energy of QL. As shown in Fig. S3 and Table S2,† QL formed different adsorption configurations on these catalysts and had different binding energies. Then, we put these alloys on the Sabatier volcano in Fig. S2.† Among these 18 alloys, NiFe and CoNi-based alloys with proper compositions have the possibility to be located at the peak of the Sabatier volcano. To further evaluate the experimental feasibility, we calculated the formation energies of these 18 alloys, which were summarized in Table S2,† where the Ni3Fe had the lowest formation energy and eventually a high stability. In this respect, we theoretically identified Ni3Fe as a good candidate with high activity and stability for the QL hydrogenation reaction.
3.2. Quinoline hydrogenation mechanism on bimetallic Ni3Fe catalyst
To further evaluate the performance of this theoretically predicted Ni3Fe catalyst, we systematically calculated the detailed mechanisms of QL hydrogenation to py-THQL. Our calculated lattice parameters of bulk Ni3Fe is a = b = c = 3.6114 Å, which agrees well with experimental values of a = b = c = 3.5525 Å.46–48 This also indicates that our DFT calculations could reasonably describe the properties of Ni3Fe alloy. As shown in Fig. 1, the hydrogenation process is very complex. For example, there are eight possibilities for the first hydrogenation step of QL on the Ni3Fe(111) surface to generate 1HQL, which directly determines the selectivity of QL hydrogenation. The eventual product formed in this process is determined by the relative ease with which H attacks a specific site of QL.
 |
| | Fig. 1 The intermediates involved in the quinoline hydrogenation mechanism over the Ni3Fe bimetallic catalyst and free energy barrier (Ga) at 353.15 K. | |
As shown in Fig. 1, the final states of 1HQL were named 1HQL-X, where the 1HQL-1, 1HQL-2, 1HQL-3, 1HQL-4, 1HQL-1′, 1HQL-2′, 1HQL-3′ and 1HQL-4′ correspond to the products of H attacking the C1, C2, C3, N4 atoms in N-heterocycle and C1′, C2′, C3′, C4′ in benzene ring, respectively. Our calculations show that the first hydrogenation will take place at the C1 site of the N-heterocycle due to the lower free energy barrier (1.03 eV) than the other possible pathways at 353.15 K. For the second hydrogenation step from 1HQL to 2HQL, we calculated the free energy barriers of seven possible hydrogenation pathways, where the second H could be added to C2, C3, N4, C1′, C2′, C3′, and C4′. Our results show that the hydrogenation of the benzene ring is less favorable than the N-heterocycle ring. Among these processes, the hydrogenation of the C2 site on the N-heterocycle ring has the lowest free energy barrier (0.88 eV). For the hydrogenation of 2HQL to 3HQL, six possible hydrogenation pathways were considered, where the hydrogenation of the C3 site on the N-heterocycle ring had the lowest free energy barrier of 0.89 eV. The final hydrogenation step occurs at the N4 site on the N-heterocycle ring with a free energy barrier of 1.05 eV. Noteworthy, the py-THQL product can not form chemical adsorption on the Ni3Fe(111) surface and is readily away from the surface (Fig. S4†), which hampers the further hydrogenation of the benzene ring to form the fully hydrogenated decahydroquinoline (DHQL).
3.3. Potential free energy diagram and MKM analysis
Based on the systematic calculations, we obtain the favorable reaction pathway of QL hydrogenation to py-THQL on the Ni3Fe catalyst, which is the successive hydrogenation of the C1, C2, C3, and N4 sites of the N-heterocycle to generate intermediates 1HQL, 2HQL, 3HQL, and py-THQL. Then, we plotted the potential free energy diagram of QL hydrogenation on the Ni3Fe(111) surface at 353.15 K in Fig. 2. It shows that the maximum free energy barrier of the most favorable reaction pathway is 1.05 eV. We calculated the quinoline hydrogenation on AuPd3 by repeating the reported pathway in our previous joint work using the same calculation parameters in this work. Indeed, these energy barriers are lower than those on our previously reported AuPd3 catalyst32 (Fig. S5†), where the free energy barrier for the rate-determining step on the AuPd3(111) surface is 1.37 eV. This indicates the potentially high activity of the predicted Ni3Fe catalyst in the selective hydrogenation of QL to py-THQL.
 |
| | Fig. 2 The potential free energy diagram of QL hydrogenation on the bimetallic catalyst Ni3Fe(111) surface at 353.15 K with free energy corrections. | |
To further evaluate the catalytic performance of the Ni3Fe(111) surface in QL hydrogenation, we established a mean-field microkinetic model to calculate the TOF of py-THQL production and more details could be found in ESI.† For convenience and clarity, we compared the activity of the Ni3Fe(111) surface with the AuPd3(111) surface. As shown in Fig. 3A and B, the TOF for py-THQL production on the Ni3Fe(111) surface is much higher than that on the AuPd3(111) surface at a wide range of reaction conditions. More specifically, at specific temperature of 500 K for quinoline hydrogenation in Fig. 3C, the theoretically calculated TOF of py-THQL production on Ni3Fe (2.29 × 10−9 s−1) is higher than AuPd3 (7.5 × 10−12 s−1) at 1 bar. Likewise, at a pressure of 100 bar in Fig. 3D, the theoretically calculated TOF is 1.86 × 10−4 s−1, which is much higher than AuPd3 (4.95 × 10−7 s−1) at a temperature of 600 K. All these computational results indicate the high potential of this theoretically predicted Ni3Fe catalyst in conducting effective QL hydrogenation.
 |
| | Fig. 3 Calculated TOF of py-THQL production as a function of temperature and total pressure at QL/H2 = 1/2 on the Ni3Fe(111) surface (A) and AuPd3(111) surface (B), as well as the pressure dependence at 500 K (C) and temperature dependence at 100 bar (D) with QL/H2 = 1/2. | |
4. Conclusions
In summary, we computationally identified Ni3Fe as a promising catalyst for the quinoline selective hydrogenation to 1,2,3,4-tetrahydroquinoline using density functional theory calculations and microkinetic modeling. Due to the high complexity of this reaction, we first established a Sabatier volcano for QL hydrogenation to py-THQL using the binding energy of QL as an adequate descriptor. Then, a systematic evaluation of the descriptor values and stabilities on 18 bimetallic alloys composed of earth-abundant elements results in identifying Ni3Fe as a good candidate with moderate binding strength of QL and high stability. Further systematic mechanism calculations on the Ni3Fe(111) surface showed that the reaction favorably followed subsequent hydrogenation of the C1, C2, C3, and N4 sites on the N-heterocycle of QL, where the energy barrier for the rate-determining step was much lower than that on the AuPd3(111) surface. Eventually, our MKM results also show a much higher TOF of this theoretically predicted Ni3Fe than the AuPd3 catalyst for py-THQL production. This work not only computationally identified an earth-abundant catalyst with high performance in QL hydrogenation but also illustrated the high capability of the Sabatier principle to guide rational catalyst design for complex reactions.
Data availability
The data supporting this article have been included as part of the ESI.†
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
T. W. is thankful for the start-up packages from Westlake University and the Research Center for Industries of the Future (RCIF) at Westlake University for supporting this work. We thank the Westlake University HPC Center for computation support.
Notes and references
- V. Sridharan, P. A. Suryavanshi and J. C. Menendez, Chem. Rev., 2011, 111, 7157–7259 CrossRef CAS PubMed.
- I. Muthukrishnan, V. Sridharan and J. C. Menendez, Chem. Rev., 2019, 119, 5057–5191 CrossRef CAS PubMed.
- Z.-Z. Wei, F.-J. Shao and J.-G. Wang, Chin. J. Catal., 2019, 40, 980–1002 CrossRef CAS.
- A. R. Katritzky, S. Rachwal and B. Rachwal, Tetrahedron, 1996, 52, 15031–15070 CrossRef CAS.
- R.-K. Chen, X.-C. Yang, H.-N. Tian and L.-C. Sun, J. Photochem. Photobiol., A, 2007, 189, 295–300 CrossRef CAS.
- Y. Hao, X.-C. Yang, J.-Y. Cong, A. Hagfeldt and L.-C. Sun, Tetrahedron, 2012, 68, 552–558 CrossRef CAS.
- K. Rakstys, J. Solovjova, T. Malinauskas, I. Bruder, R. Send, A. Sackus, R. Sens and V. Getautis, Dyes Pigm., 2014, 104, 211–219 CrossRef CAS.
- T. Kubo, C. Katoh, K. Yamada, K. Okano, H. Tokuyama and T. Fukuyama, Tetrahedron, 2008, 64, 11230–11236 CrossRef CAS.
- K. Maruoka, T. Miyazaki, M. Ando, Y. Matsumura, S. Sakane, K. Hattori and H. Yamamoto, J. Am. Chem. Soc., 2002, 105, 2831–2843 CrossRef.
- R. H. Fish, A. D. Thormodsen and G. A. Cremer, J. Am. Chem. Soc., 1982, 104, 5234–5237 CrossRef CAS.
- M. Campanati, A. Vaccari and O. Piccolo, J. Mol. Catal. A: Chem., 2002, 179, 287–292 CrossRef CAS.
- J. Xia, D. Wang, R. Yang, Y. Deng and G.-J. Deng, Green Chem., 2024, 26, 5160–5166 RSC.
- M. Wang, C. Zhang, C. Ci, H. Jiang, P. H. Dixneuf and M. Zhang, J. Am. Chem. Soc., 2023, 145, 10967–10973 CrossRef CAS PubMed.
- S. Zhang, Z.-M. Xia, T. Ni, H. Zhang, C. Wu and Y.-Q. Qu, J. Mater. Chem. A, 2017, 5, 3260–3266 RSC.
- K. Li, B.-C. Hao, M. Xiao, Y.-W. Kuang, H.-Y. Shang, J. Ma, Y. Liao and H. Mao, Appl. Surf. Sci., 2019, 478, 176–182 CrossRef CAS.
- G. Perot, Catal. Today, 1991, 10, 447–472 CrossRef CAS.
- L. Tao, Q. Zhang, S.-S. Li, X. Liu, Y.-M. Liu and Y. Cao, Adv. Synth. Catal., 2015, 357, 753–760 CrossRef CAS.
- X.-F. Tu and L.-Z. Gong, Angew. Chem., Int. Ed., 2012, 51, 11346–11349 CrossRef CAS PubMed.
- S. Murahashi, Y. Imada and Y. Hirai, Bull. Chem. Soc. Jpn., 1989, 62, 2968–2976 CrossRef CAS.
- G. E. Dobereiner, A. Nova, N. D. Schley, N. Hazari, S. J. Miller, O. Eisenstein and R. H. Crabtree, J. Am. Chem. Soc., 2011, 133, 7547–7562 CrossRef CAS PubMed.
- Y.-T. Gong, P.-F. Zhang, X. Xu, Y. Li, H.-R. Li and Y. Wang, J. Catal., 2013, 297, 272–280 CrossRef CAS.
- S. Zhang, Z.-M. Xia, T. Ni, Z.-Y. Zhang, Y.-Y. Ma and Y.-Q. Qu, J. Catal., 2018, 359, 101–111 CrossRef CAS.
- M. Fang and R. A. Sánchez-Delgado, J. Catal., 2014, 311, 357–368 CrossRef CAS.
- A. Sánchez, M.-F. Fang, A. Ahmed and R. A. Sánchez-Delgado, Appl. Catal., A, 2014, 477, 117–124 CrossRef.
- Y. Hu, X. Li, M. Liu, S. Bartling, H. Lund, J. Rabeah, P. J. Dyson, M. Beller and R. V. Jagadeesh, ChemCatChem, 2024, 16, e202301027 CrossRef CAS.
- X. Wei, Y. Jiang, Y. Ma, H. Liao, S. Dai, P. An, Z.-Q. Wang, X.-Q. Gong and Z. Hou, ACS Catal., 2024, 14, 5344–5355 CrossRef CAS.
- J.-W. Zhang, D.-D. Li, G.-P. Lu, T. Deng and C. Cai, ChemCatChem, 2018, 10, 4966–4972 CrossRef.
- Y.-G. Chen, Z.-J. Yu, Z. Chen, R.-G. Shen, Y. Wang, X. Cao, Q. Peng and Y.-D. Li, Nano Res., 2016, 9, 2632–2640 CrossRef CAS.
- J.-Q. Bai, M. Tamura, Y. Nakagawa and K. Tomishige, Catal. Sci. Technol., 2022, 12, 2420–2425 RSC.
- R.-R. Yun, L.-R. Hong, W.-J. Ma, S.-N. Wang and B.-S. Zheng, ACS Appl. Nano Mater., 2019, 2, 6763–6768 CrossRef CAS.
- J.-B. Zhao, H.-F. Yuan, G. Yang, Y.-F. Liu, X.-M. Qin, Z. Chen, C. Weng-Chon, L.-M. Zhou and S.-M. Fang, Nano Res., 2021, 15, 1796–1802 CrossRef.
- X.-J. Cui, Z.-J. Huang, A. P. van Muyden, Z.-F. Fei, T. Wang and P. J. Dyson, Sci. Adv., 2020, 6, eabb3831 CrossRef CAS PubMed.
- P. Zheng, C. Xiao, S. Song, A. Duan and C. Xu, J. Hazard. Mater., 2021, 411, 125127 CrossRef CAS PubMed.
- L.-F. Yin, J.-L. Ma, L.-X. Ling, R.-G. Zhang, G.-C. Yan, J.-L. Wang, W.-M. Lu, Y. Li and B.-J. Wang, Mol. Catal., 2023, 538, 112970 CrossRef CAS.
- C. J. Jacobsen, S. Dahl, B. S. Clausen, S. Bahn, A. Logadottir and J. K. Norskov, J. Am. Chem. Soc., 2001, 123, 8404–8405 CrossRef CAS PubMed.
- J. K. Nørskov, T. Bligaard, J. Rossmeisl and C. H. Christensen, Nat. Chem., 2009, 1, 37–46 CrossRef PubMed.
- A. J. Medford, A. Vojvodic, J. S. Hummelshøj, J. Voss, F. Abild-Pedersen, F. Studt, T. Bligaard, A. Nilsson and J. K. Nørskov, J. Catal., 2015, 328, 36–42 CrossRef CAS.
- S.-H. Liu, N. Govindarajan and K. Chan, ACS Catal., 2022, 12, 12902–12910 CrossRef CAS.
- T. Wang, J. Ibañez, K. Wang, L. Fang, M. Sabbe, C. Michel, S. Paul, M. Pera-Titus and P. Sautet, Nat. Catal., 2019, 2, 773–779 CrossRef CAS.
- G. M. Xu, C. Cai and T. Wang, J. Am. Chem. Soc., 2022, 144, 23089–23095 CrossRef CAS PubMed.
- G. M. Xu, C. Cai, W. H. Zhao, Y. H. Liu and T. Wang, WIREs Comput. Mol. Sci., 2022, 13, e1654 CrossRef.
- G. Kresse and J. Furthmüller, Comput. Mater. Sci., 1996, 6, 15–50 CrossRef CAS.
- J. Wellendorff, K. T. Lundgaard, A. Møgelhøj, V. Petzold, D. D. Landis, J. K. Nørskov, T. Bligaard and K. W. Jacobsen, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 235149 CrossRef.
- G. Henkelman, B. P. Uberuaga and H. Jónsson, J. Chem. Phys., 2000, 113, 9901–9904 CrossRef CAS.
- S. R. Bahn and K. W. Jacobsen, Comput. Sci. Eng., 2002, 4, 56–66 CrossRef CAS.
- P. Paufler, Cryst. Res. Technol., 1987, 22, 1436 CrossRef.
- I. Chicinas, V. Pop and O. Isnard, J. Magn. Magn. Mater., 2002, 242–245, 885–887 CrossRef CAS.
- I. Chicinaş, V. Pop, O. Isnard, J. M. Le Breton and J. Juraszek, J. Alloys Compd., 2003, 352, 34–40 CrossRef.
|
| This journal is © The Royal Society of Chemistry 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.