Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

This article is licensed under CC-BY 4.

pubs.acs.org/orginorgau Article

Biomimetic Frustrated Lewis Pair Catalysts for Hydrogenation of CO


to Methanol at Low Temperatures
Jiejing Zhang, Longfei Li,* Xiaofeng Xie, Xue-Qing Song,* and Henry F. Schaefer, III*
Cite This: https://doi.org/10.1021/acsorginorgau.3c00064 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: The industrial production of methanol through CO hydro-


Downloaded via INDIAN INST OF TECH PALAKKAD on February 10, 2024 at 12:52:14 (UTC).

genation using the Cu/ZnO/Al2O3 catalyst requires harsh conditions, and the
development of new catalysts with low operating temperatures is highly
desirable. In this study, organic biomimetic FLP catalysts with good tolerance
to CO poison are theoretically designed. The base-free catalytic reaction
contains the 1,1-addition of CO into a formic acid intermediate and the
hydrogenation of the formic acid intermediate into methanol. Low-energy
spans (25.6, 22.1, and 20.6 kcal/mol) are achieved, indicating that CO can be
hydrogenated into methanol at low temperatures. The new extended
aromatization−dearomatization effect involving multiple rings is proposed
to effectively facilitate the rate-determining CO 1,1-addition step, and a new
CO activation model is proposed for organic catalysts.
KEYWORDS: CO hydrogenation, biomimetic FLP catalysts, density functional theory, 1,1-addition, base-free

1. INTRODUCTION to methanol catalyzed by various transition-metal carbonyls


Alternative energy sources have received great attention due to has been reported, but the problem of extreme reaction
the limited supply and increasing cost of fossil fuel resources.1,2 conditions was not solved.14−20 In general, CO is a high-field
Methanol is an excellent fuel in its own right and can also ligand, and its migratory insertion into a metal−hydride bond
produce energy in electrochemical cells or other energy vectors is highly endothermic, which is responsible for the harsh
such as hydrogen.3−5 Besides, methanol is a key feedstock for reaction conditions.21,22 Instead, avoiding this highly endo-
the production of small-chain hydrocarbons.6,7 Olah and co- thermic CO insertion step could allow mild reaction
conditions for the synthesis of methanol.23
workers previously advocated “methanol economy” as a way
Recently, bifunctional metal catalysts have been reported to
toward the sustainable economy, in which petroleum-based
catalyze the hydrogenolysis of esters, amides, and their
fuels and chemicals are replaced with methanol.8 However,
derivatives to produce alcohols.24−29 Those research studies
methanol is industrially produced from CO hydrogenation
inspired the design of bifunctional metal catalysts for the
using Cu/ZnO/Al2O3 heterogeneous catalysts at elevated
indirect hydrogenation of CO to methanol via ester or amide
pressures (50−100 bar) and temperatures (200−300 °C).9,10
intermediates. In 2019, Prakash and co-workers found that
Since the hydrogenation of CO to methanol is an exothermic
electron-withdrawing phenyl groups could render the Ru−
reversible reaction, CO conversion in current processes is
Macho−BH complex resistant to CO poisoning and achieved
thermodynamically limited to around 30% at such high
the Ru−Macho−BH catalyzed indirect hydrogenation of CO
operating temperatures.11 Therefore, the development of
to methanol at 145 °C with a turnover number (TON) of 539
homogeneous catalysts for CO hydrogenation at low temper-
(Scheme 2a).12 Prakash also proposed that CO was first
atures would be a significant progress toward increasing the
anchored onto the amine as a formamide and then hydro-
energy efficiency and CO conversions.12 The main challenges
genated in situ to methanol. In the same year, Beller and co-
are the high endothermic migratory insertion process of CO
workers investigated a series of metal complexes for the
into metal hydride bonds and the poisons of CO to metal
indirect hydrogenation of CO to methanol through N-
catalysts (Scheme 1).
The reported homogeneous CO hydrogenations rely on
metal catalysts, and the mechanisms can be broadly divided Received: November 27, 2023
into direct and indirect mechanisms. In the early 1950s, Revised: January 12, 2024
DuPont Company first reported the direct hydrogenation of Accepted: January 16, 2024
CO to methanol catalyzed by cobalt carbonyl complexes, but
the reaction required ultrahigh pressures (1500−5000 atm).13
In the following three decades, the direct hydrogenation of CO
© XXXX The Authors. Published by
American Chemical Society https://doi.org/10.1021/acsorginorgau.3c00064
A ACS Org. Inorg. Au XXXX, XXX, XXX−XXX
ACS Organic & Inorganic Au pubs.acs.org/orginorgau Article

Scheme 1. Catalysts for the Hydrogenation of CO to Methanol

Scheme 2. Reported Bifunctional Metal Catalysts for the Indirect Hydrogenation of CO to Methanol

formylazole intermediates, and the manganese pincer complex suitable π-back-bonding orbitals.35 This implies good tolerance
exhibited the best tolerance to CO and gave the optimal TON to CO and inspires us to explore main-group element catalysts
(3170) at 120−150 °C (Scheme 2b).30 In 2021, Leitner and for the indirect hydrogenation of CO to methanol. Frustrated
co-workers reported an alcohol-assisted indirect hydrogenation Lewis pairs (FLPs) introduced by Stephan and co-workers are
of CO to methanol catalyzed by a manganese pincer complex now an important branch of main-group element chemis-
with a catalytic amount of base at 150 °C (Scheme 2c).31 They try36,37 and have been applied in the catalytic hydrogenation of
found that the increased pressure of CO could inhibit various unsaturated compounds including CO2,38 olefins,39
methanol formation. These remarkable bifunctional metal alkynes,40 ketones,41 and imines.42 Particularly, the FLP-
catalysts facilitate indirect CO hydrogenation but are still catalyzed hydrogenation of amides43 and esters44,45 to provide
limited by high reaction temperatures and low yields, which alcohols was also reported recently. However, FLP-catalyzed
could be associated with the deactivation of transition-metal CO hydrogenation has not been reported yet. How can we
catalysts in the presence of CO.32 endow FLP catalysts with high activities toward CO hydro-
Main-group element catalysts are desirable alternatives to genations?
transition-metal catalysts because of their natural abundance The design of novel bioinspired catalysts mimicking the
and low cost.33,34 More importantly, main-group element mechanism of biocatalysts is an important strategy to obtain
catalysts have low binding affinity to CO, owing to the lack of high catalytic activity, yield, and selectivity.46−49 The nucleic
B https://doi.org/10.1021/acsorginorgau.3c00064
ACS Org. Inorg. Au XXXX, XXX, XXX−XXX
ACS Organic & Inorganic Au pubs.acs.org/orginorgau Article

Scheme 3. Nucleic Acid Bases and Bioinspired FLP Catalysts

acid bases (Scheme 3) with resonance stabilization effects can


undergo protonation or tautomerization50−53 and play an
important role in the reproduction and transmission of genetic
information.54 Can the Lewis basic nucleic acid bases form
novel FLP catalysts with borane counterparts and achieve new
chemical reactivities? The theoretical design plays a more and
more important role in discovering new chemistry, and many
achievements have been made to solve important problems in
chemistry.55−60 Inspired by our and other groups’ silico
reaction discoveries,61−67 we theoretically design bioinspired
FLP catalysts based on nucleic acid bases and compare their
catalytic activities in the CO hydrogenation reaction to
methanol with the traditional FLP catalysts.

2. COMPUTATIONAL METHODS Figure 1. Structure of bioinspired FLP A1 and the catalyzed


In accord with our previous theoretical studies,68−70 this research was hydrogenation of CO to methanol.
carried out with the DFT ωB97X-D71 method using the Gaussian 09
program.72 Geometries were optimized in a dichloromethane solvent 3.1. 1,1-Addition of CO to the Formic Acid Intermediate
using the 6-311G(d,p)73 basis sets. The single-point energy
refinements were further performed with the 6-311++G(2d,p) basis
The bioinspired FLP A1 features good tolerance to CO, and
sets. The refined energies were then corrected to Gibbs energies at the coordination of CO to A1 will afford a thermodynamically
298.15 K and 1 atm by using the ωB97X-D/6-311G(d,p) harmonic unstable intermediate A1_CO, being endergonic by 5.2 kcal/
frequencies. Solvent effects were evaluated using the SMD (solution mol as shown in Figure 2. The dimerization of A1 through
model based on density) solvation model.74,75 All transition states forming a Lewis adduct is thermodynamically disfavored, being
were demonstrated to exhibit only one imaginary vibrational endergonic by 2.4 kcal/mol (Figure S2), suggesting that the
frequency. Intrinsic reaction coordinate (IRC) analyses were monomer of A1 is the major species in solution. The B atom of
performed to confirm that all transition states connect the two A1 is Lewis acidic with an NBO charge of 0.94 and can bind a
minima in question.76 Natural bond orbital (NBO) analyses were H 2 O molecule through the O → B donor−acceptor
performed using the NBO-7.0 program.77 The Cartesian coordinates interaction, being endergonic by 4.3 kcal/mol. Through the
of all optimized structures are presented in the Supporting transition state TSA2−3, the O−H bond of the coordinated
Information. H2O molecule is cleaved, and the hydroxy proton H(2) moves
to the adjacent carbonylic O atom with an energy barrier of 8.5
3. RESULTS AND DISCUSSION kcal/mol. Subsequently, the zwitterionic proton−hydroxo
intermediate A3 is formed, in which the C(1)−N(1) and
The bioinspired FLP A1 featuring a nucleic acid base-fused π- C(1)−N(2) bonds are significantly shortened compared with
conjugated linker is designed, and the corresponding structure those in A1. The nucleus-independent chemical shift (NICS)
is shown in Figure 1. The distance between the Lewis acidic B method is widely utilized for determining aromaticity.78 The
atom and the Lewis basic carbonylic O atom is large (5.34 Å), NICS value of the pyrimidinol ring in A3 (ring 1; −6.49 ppm)
which provides a suitable pocket for the synchronous is much more negative than that of the pyrimidinone ring in
activation of two substrate molecules (e.g., water and CO). A1 (ring 1; −2.80 ppm), suggesting that the pyrimidinol ring
The A1-catalyzed CO hydrogenation to methanol is designed in A3 has become aromatized. As for ring 2 and ring 4 in A3,
and is composed of two stages: (1) the 1,1-addition of CO into the NICS values become more negative by 1.54 and 1.20 ppm,
formic acid and (2) the hydrogenation of formic acid into respectively; therefore, the protonation of the carbonylic O
methanol. atom also renders the remote ring 2 and ring 4 more aromatic.
C https://doi.org/10.1021/acsorginorgau.3c00064
ACS Org. Inorg. Au XXXX, XXX, XXX−XXX
ACS Organic & Inorganic Au pubs.acs.org/orginorgau Article

Figure 2. Gibbs energy profile for the 1,1-addition of CO to a formic acid intermediate catalyzed by the bioinspired FLP A1. The relative Gibbs
energies (ΔG) and potential energies (ΔE) are in kcal/mol.

Figure 3. Atomic charges, bond lengths, and the nucleus-independent chemical shift (NICS) values of the rings in A1, A3, and A7. The −tBu and
−C6F5 groups are drawn in a wireframe for simplicity.

The NICS value of ring 3 in A3 increases a bit (0.27 ppm) carbonylic C atom of CO, simultaneously. The 1,1-addition of
(Figure 3). The extended aromatization−dearomatization CO also involves the extended aromatization−dearomatization
effect involving multiple rings was not reported previously. effect, and intermediate A4 with a dearomatized pyrimidinone
The proton and hydroxy moieties in zwitterionic A3 can be ring is formed. The following dissociation of the formic acid
added to CO through an unusual 1,1-addition process. The molecule from A4 will regenerate catalyst A1, and the stage of
H(2)...O(2) distance in A3 is 3.81 Å, and the O(2) atom is CO 1,1-addition is achieved. Including the energy profile of
nucleophilic with an NBO charge of −0.95. After the transition Stage 2 included in Section 3.2, the 1,1-addition of CO via
state TSA3−4, proton H(2) and hydroxo transfer to the TSA3−4 is the rate-determining step of the overall catalytic
D https://doi.org/10.1021/acsorginorgau.3c00064
ACS Org. Inorg. Au XXXX, XXX, XXX−XXX
ACS Organic & Inorganic Au pubs.acs.org/orginorgau Article

Figure 4. Comparisons between the typical traditional FLP catalysts and the new bioinspired FLP catalysts via the Gibbs energy barriers (ΔG‡) of
the rate-determining CO 1,1-addition.

Figure 5. (a) Dewar−Chatt−Duncanson model for CO activation based on metal catalysts; (b) dynamic natural localized molecular orbital
(NLMO) analysis for the unusual CO 1,1-addition process; and (c) new CO activation model based on organic catalysts.

reaction with an energy span of 25.6 kcal/mol. Other The acceleration effects of the extended aromatization−
promoters are also considered for the 1,1-addition of CO in dearomatization of our designed bioinspired FLP catalysts are
Figure S3 in SI, and the methanol promoter under catalyst A1 studied. The typical traditional catalysts and the designed
is predicted to own a high-energy barrier of 28.1 kcal/mol. bioinspired FLP catalysts are compared to the Gibbs energy
E https://doi.org/10.1021/acsorginorgau.3c00064
ACS Org. Inorg. Au XXXX, XXX, XXX−XXX
ACS Organic & Inorganic Au pubs.acs.org/orginorgau Article

Figure 6. Hydrogen shuttle mechanism for hydrogen activation by bioinspired FLP A1. The relative Gibbs energies (ΔG) and potential energies
(ΔE) are in kcal/mol.

barriers (ΔG‡) of the rate-determining CO 1,1-addition steps contrast, the CO activation by organic catalysts does not
via TSA3−4. The corresponding whole catalytic pathways are involve dπ-back-donation, and it is still unclear. In order to
shown in Figures S6−S17 in the SI. The typical traditional FLP unveil the reasons for CO activation by the bioinspired FLP,
catalysts including intermolecular and intramolecular systems the unusual 1,1-addition of CO was simulated. This was
are predicted to involve high-energy barriers (34−46 kcal/mol; accomplished via performing the intrinsic reaction coordinate
Figure 4a). It is worth noting that the traditional FLPs H1 and (IRC) computation on the transition state TSA3−4. The
I1 have the similar linker structure with the designed natural localized molecular orbital (NLMO) analysis was
bioinspired FLP A1, and the distances of Lewis acid−base applied to five frames of the IRC trajectory, as shown in Figure
sites are almost the same for H1, I1, and A1 (ca. 5.3 Å). 5b. When the CO molecule approaches A3, the lone pair of the
Differently, H1 and I1 cannot exhibit the extended carbonyl carbon atom is delocalized to the σ*O−H antibonding
aromatization−dearomatization effect. Consequently, the orbital, and the lone pair electron of the O(2) atom is
energy barriers of H1 and I1 catalysts (38.3 and 34.5 kcal/ delocalized to the π*C�O orbital. According to our dynamic
mol) are predicted to be much higher than that of the NLMO analysis, a new CO activation model involving the σ-
bioinspired FLP A1 catalysts (25.6 kcal/mol), suggesting that donation and σ-back-donation interactions is proposed, as
the extended aromatization−dearomatization in A1 could play illustrated in Figure 5c. This, hopefully, will guide the
an important role in facilitating the CO 1,1-addition step. development of organic catalysts for CO hydrogenations.
Further, the activities of the bioinspired FLP catalysts can be 3.2. Hydrogenation of Formic Acid Intermediate into
controlled. As shown in Figure 4b, the intermolecular catalyst Methanol
J1 has a higher energy barrier of the rate-determining CO 1,1- The Lewis basic O and acidic B atoms in A1 cannot form the
addition steps than the intramolecular ones, which could be effective orbital overlap with the hydrogen molecule due to the
associated with the entropy penalty. The intramolecular long O...B distance (5.34 Å) and make the common hydrogen
bioinspired FLP catalyst K1 with −Ph groups is predicted to activation pathway difficult. As shown in Figure 6, the
own an energy barrier (31.8 kcal/mol) higher than that of A1 hydrogen molecule cleaves across the B and O atoms of A1
with −B(C6F5)2 groups (25.6 kcal/mol), suggesting that the with a high energy barrier (39.1 kcal/mol), due to the high
higher Lewis acidity of B atom will result in lower energy strain of the hydrogen-transfer transition state TSA1−7,
barriers. Pursuing this idea, we introduced more F substituents indicating that this pathway is not available. Instead, a
on the linkers, and the modified FLP catalysts L1 and M1 are hydrogen shuttle mechanism becomes available for the
predicted to yield even lower energy barriers (22.1 and 20.6 bioinspired FLP-facilitated hydrogen activation. The water
kcal/mol), which can allow for low operating temperatures. and hydrogen molecules can approach A1 to form the complex
The activation of CO by metal complexes is a highly A5, and this step is endothermic by 23.7 kcal/mol due to the
important field in chemistry, and it has been explained in terms entropy penalty. Through transition state TSA5−6, the H(3)
of the Dewar−Chatt−Duncanson model, which involves the σ- atom moves from the hydrogen moiety to the nearby O(2)
donation and dπ-back-donation interactions (Figure 5a).79 In atom of the water moiety, which is accompanied by the H(2)
F https://doi.org/10.1021/acsorginorgau.3c00064
ACS Org. Inorg. Au XXXX, XXX, XXX−XXX
ACS Organic & Inorganic Au pubs.acs.org/orginorgau Article

Figure 7. Gibbs energy profile for the hydrogenation of formic acid to methanol catalyzed by the bioinspired FLP. The relative Gibbs energies
(ΔG) and potential energies (ΔE) are given in kcal/mol.

atom transfer from the water moiety to the carbonylic O(1) own even lower energy spans (22.1 and 20.6 kcal/mol),
atom. Actually, the water molecule serves as a hydrogen shuttle suggesting that the designed CO hydrogenation to methanol
and decreases the energy barrier of the hydrogen activation to could be achieved at room temperatures according to both
23.3 kcal/mol. Note that although the Gibbs energy of TSA5− reported theoretical studies80−84 and Eyring equation that
6 (21.0 kcal/mol) appears to be lower than that of A5 (21.4 correlates the activation energy (ΔG‡) and the rate constant
kcal/mol), the potential energy of TSA5−6 (−15.0 kcal/mol) (k) of the rate-determining step.85
is higher than that of A5 (−15.7 kcal/mol). In the following,
the intermediate A6 is generated and can release a water 4. CONCLUSIONS
molecule to form the hydrogenated intermediate A7. As shown
in Figure 3, the bond lengths and NICS values of A7 are The organic bioinspired FLP A1 featuring a nucleic acid base
almost the same with those in A3, suggesting that the hydrogen fused π-conjugated linker is theoretically designed and can
activation also experiences the extended aromatization− provide a suitable pocket for the synchronous activation of two
dearomatization process. substrate molecules. The computations suggest that the
As shown in Figure 7, the H(2)...H(4) distance in the biomimetic FLP catalyst can exhibit good tolerance to CO
structure of A7 is 3.84 Å, and the NBO charges for H(2) and poison. The base-free water-assisted indirect CO hydro-
H(4) atoms are 0.52 and −0.02, respectively. Through the genation to methanol catalyzed by the bioinspired FLP
transition state TSA7, the H(2) and H(4) atoms can transfer catalysts contains two stages: the 1,1-addition of CO into the
to the C�O bond of formic acid in a concerted manner with formic acid intermediate and the hydrogenation of the formic
an energy barrier of 16.5 kcal/mol. As a result, methane diol acid intermediate into methanol. The 1,1-addition of CO via
and the catalyst A1 are formed, being exoenergic by 5.0 kcal/ the transition state TSA3−4 is the rate-determining step with
mol. In the following, FLP A1 can facilitate the decomposition an energy span of 25.6 kcal/mol. Furthermore, the activities of
of methane diol to form formaldehyde. Through the H...O the bioinspired FLP catalysts can be controlled through
hydrogen-bonding and O→B coordinative interactions, A1 can changing the Lewis acidity of the B atom, and the modified
bind methane diol to afford the intermediate A8. Over the catalysts L1 and M1 yield lower energy spans (22.1 and 20.6
transition state TSA8, the methane diol moiety in A8 can be
kcal/mol). These low-energy spans suggest that CO can be
decomposed with the cleavage of C(3)−O(2) and O(3)−
hydrogenated into methanol at low operating temperatures.
H(2) bonds, and the corresponding energy barrier is only 6.0
The comparisons between the typical traditional catalysts
kcal/mol. After that, formaldehyde is released. The generated
intermediate A3 will release a water molecule through the and the designed bioinspired FLP catalysts unveil that the new
water activation pathway in Figure 2 (A3 → TSA2−3 → A2 extended aromatization−dearomatization effect involving
→ A1). The bioinspired FLP A1 then undergoes hydrogen multiple rings of bioinspired FLP catalysts is an effective
activation (see pathway in Figure 6) and the C�O bond strategy to facilitate the rate-determining CO 1,1-addition step.
hydrogenation process and forms the final methanol product. According to the dynamic NLMO analysis, a new model
In brief, the reaction heat of the whole reaction is −18.2 involving σ-donation and σ-back-donation interactions is
kcal/mol, and the 1,1-addition of CO via the transition state proposed for CO activation and should guide the development
TSA3−4 (stage 1) is the rate-determining step with an energy of organic catalysts for CO hydrogenations. The present study
span of 25.6 kcal/mol for the bioinspired FLP A1. Moreover, will pave the way to a new method for the hydrogenation of
the pathways created by the modified bioinspired FLPs L1 and CO to methanol with good stability, low cost, environmental
M1 are shown in Figures S16 and S17 in SI and predicted to friendliness, and low temperatures.
G https://doi.org/10.1021/acsorginorgau.3c00064
ACS Org. Inorg. Au XXXX, XXX, XXX−XXX
ACS Organic & Inorganic Au pubs.acs.org/orginorgau Article

■ ASSOCIATED CONTENT
Data Availability Statement
■ ACKNOWLEDGMENTS
This work was supported by the National Natural Science
The data underlying this study are available in the published Foundation of China (No. 22373030), the Natural Science
article and its Supporting Information. Interdisciplinary Research Program of Hebei University (No.
DXK202013), the Jointed Key Pharmaceutical and Drug
*
sı Supporting Information
Project of Hebei Province (B2022208057), and the Innovation
The Supporting Information is available free of charge at Capacity Improvement Plan of Hebei Province (20567605H).
https://pubs.acs.org/doi/10.1021/acsorginorgau.3c00064. The authors thank the High Performance Computer Center of
Hebei University for providing computational resources. HFS
Optimized geometries and energies of all stationary was supported by the U.S. Department of Energy, under
points along the reaction pathways and the imaginary contract No. DE-SC0018412.
vibrational frequencies of transition states (PDF)

■ AUTHOR INFORMATION
■ REFERENCES
(1) Turner, J. M. The matter of a clean energy future. Science 2022,
Corresponding Authors 376, 1361.
(2) Slameršak, A.; Kallis, G.; O’Neill, D. W. Energy requirements
Longfei Li − College of Pharmacy, Key Laboratory of and carbon emissions for a low-carbon energy transition. Nat.
Pharmaceutical Quality Control of Hebei Province, Key Commun. 2022, 13, 6932.
Laboratory of Medicinal Chemistry and Molecular Diagnosis (3) Sen, R.; Goeppert, A.; Kar, S.; Prakash, G. K. S. Hydroxide Based
of Ministry of Education, Hebei University, Baoding 071002 Integrated CO2 Capture from Air and Conversion to Methanol. J. Am.
Hebei, P. R. China; orcid.org/0000-0001-8435-4006; Chem. Soc. 2020, 142, 4544−4549.
Email: lilongfei@hbu.edu.cn (4) Zhang, Z.; Mao, C.; Meira, D. M.; Duchesne, P. N.; Tountas, A.
Xue-Qing Song − College of Pharmacy, Key Laboratory of A.; Li, Z.; Qiu, C.; Tang, S.; Song, R.; Ding, X.; Sun, J.; Yu, J.; Howe,
Pharmaceutical Quality Control of Hebei Province, Key J. Y.; Tu, W.; Wang, L.; Ozin, G. A. New black indium oxide-tandem
photothermal CO2-H2 methanol selective catalyst. Nat. Commun.
Laboratory of Medicinal Chemistry and Molecular Diagnosis
2022, 13, 1512.
of Ministry of Education, Hebei University, Baoding 071002 (5) Amann, P.; Klötzer, B.; Degerman, D.; Köpfle, N.; Götsch, T.;
Hebei, P. R. China; orcid.org/0000-0001-5527-5261; Lömker, P.; Rameshan, C.; Ploner, K.; Bikaljevic, D.; Wang, H. Y.;
Email: sxqing@hbu.edu.cn Soldemo, M.; Shipilin, M.; Goodwin, C. M.; Gladh, J.; Halldin
Henry F. Schaefer, III − Center for Computational Quantum Stenlid, J.; Börner, M.; Schlueter, C.; Nilsson, A. The state of zinc in
Chemistry, University of Georgia, Athens, Georgia 30602, methanol synthesis over a Zn/ZnO/Cu(211) model catalyst. Science
United States; orcid.org/0000-0003-0252-2083; 2022, 376, 603−608.
Email: ccq@uga.edu (6) Zhang, F.; Chen, K.; Jiang, Q.; He, S.; Chen, Q.; Liu, Z.; Kang,
J.; Zhang, Q.; Wang, Y. Selective Transformation of Methanol to
Authors Ethanol in the Presence of Syngas over Composite Catalysts. ACS
Jiejing Zhang − College of Pharmacy, Key Laboratory of Catal. 2022, 12, 8451−8461.
Pharmaceutical Quality Control of Hebei Province, Key (7) Wang, C.; Yang, L.; Gao, M.; Shao, X.; Dai, W.; Wu, G.; Guan,
N.; Xu, Z.; Ye, M.; Li, L. Directional Construction of Active
Laboratory of Medicinal Chemistry and Molecular Diagnosis
Naphthalenic Species within SAPO-34 Crystals toward More Efficient
of Ministry of Education, Hebei University, Baoding 071002
Methanol-to-Olefin Conversion. J. Am. Chem. Soc. 2022, 144, 21408−
Hebei, P. R. China 21416.
Xiaofeng Xie − College of Pharmacy, Key Laboratory of (8) Olah, G. A. Beyond Oil and Gas: The Methanol Economy.
Pharmaceutical Quality Control of Hebei Province, Key Angew. Chem., Int. Ed. 2005, 44, 2636−2639.
Laboratory of Medicinal Chemistry and Molecular Diagnosis (9) Behrens, M.; Studt, F.; Kasatkin, I.; Kühl, S.; Hävecker, M.;
of Ministry of Education, Hebei University, Baoding 071002 Abild-Pedersen, F.; Zander, S.; Girgsdies, F.; Kurr, P.; Kniep, B. L.;
Hebei, P. R. China; orcid.org/0000-0002-8113-4064 Tovar, M.; Fischer, R. W.; No̷ rskov, J. K.; Schlögl, R. The active site of
Complete contact information is available at: methanol synthesis over Cu/ZnO/Al2O3 industrial catalysts. Science
https://pubs.acs.org/10.1021/acsorginorgau.3c00064 2012, 336, 893−897.
(10) Laudenschleger, D.; Ruland, H.; Muhler, M. Identifying the
nature of the active sites in methanol synthesis over Cu/ZnO/Al2O3
Author Contributions catalysts. Nat. Commun. 2020, 11, 3898.
CRediT: Jiejing Zhang data curation, investigation, method- (11) Sen, R.; Goeppert, A.; Surya Prakash, G. K. Homogeneous
ology, software, writing-original draft; Longfei Li conceptual- Hydrogenation of CO2 and CO to Methanol: The Renaissance of
ization, funding acquisition, methodology, project adminis- Low-Temperature Catalysis in the Context of the Methanol
tration, supervision, writing-original draft, writing-review & Economy. Angew. Chem., Int. Ed. 2022, 61, No. e202207278.
(12) Kar, S.; Goeppert, A.; Prakash, G. K. S. Catalytic Homogeneous
editing; Xiaofeng Xie data curation, investigation, method- Hydrogenation of CO to Methanol via Formamide. J. Am. Chem. Soc.
ology, visualization, writing-original draft; Xue-Qing Song 2019, 141, 12518−12521.
formal analysis, investigation, validation, writing-review & (13) Walter, R. Iron catalysts for carbon monoxide hydrogenation
editing; Henry F. Schaefer conceptualization, funding precipitated from ferrous sulfate solutions. U.S. Patent 2534018, 1950.
acquisition, methodology, resources, supervision, writing- (14) Gresham, W. F. Preparation of polyfunctional compounds. U.S.
review & editing. Patent 2636046, 1953.
(15) Rathke, J. W.; Feder, H. M. Catalysis of carbon monoxide
Notes
hydrogenation by soluble mononuclear complexes. J. Am. Chem. Soc.
The authors declare no competing financial interest. 1978, 100, 3623−3625.

H https://doi.org/10.1021/acsorginorgau.3c00064
ACS Org. Inorg. Au XXXX, XXX, XXX−XXX
ACS Organic & Inorganic Au pubs.acs.org/orginorgau Article

(16) Bradley, J. S. Homogeneous carbon monoxide hydrogenation (38) Barrales-Martínez, C.; Durán, R.; Jaque, P. Metal-free catalytic
to methanol catalyzed by soluble ruthenium complexes. J. Am. Chem. conversion of CO2 into methanol: local electrophilicity as a tunable
Soc. 1979, 101, 7419−7421. property in the design and performance of aniline-derived amino-
(17) Dombek, B. D. Hydrogenation of carbon monoxide to borane-based FLPs. Inorg. Chem. Front. 2023, 10, 2344−2358.
methanol and ethylene glycol by homogeneous ruthenium catalysts. (39) Greb, L.; Oña-Burgos, P.; Schirmer, B.; Grimme, S.; Stephan,
J. Am. Chem. Soc. 1980, 102, 6855−6857. D. W.; Paradies, J. Metal-free catalytic olefin hydrogenation: low-
(18) Fahey, D. R. Rational mechanism for homogeneous hydro- temperature H2 activation by frustrated Lewis pairs. Angew. Chem., Int.
genation of carbon monoxide to alcohols, polyols, and esters. J. Am. Ed. 2012, 51, 10164−10168.
Chem. Soc. 1981, 103, 136−141. (40) Chernichenko, K.; Madarász, A.; Pápai, I.; Nieger, M.; Leskelä,
(19) Dombek, B. D. Roles of metal carbonyl complexes in the M.; Repo, T. A frustrated-Lewis-pair approach to catalytic reduction
homogeneous hydrogenation of carbon monoxide. J. Organomet. of alkynes to cis-alkenes. Nat. Chem. 2013, 5, 718−723.
Chem. 1989, 372, 151−161. (41) Mahdi, T.; Stephan, D. W. Enabling catalytic ketone
(20) Marko, L. C2 compounds directly from synthesis gas via hydrogenation by frustrated Lewis pairs. J. Am. Chem. Soc. 2014,
organometallic catalysts. Transition Met. Chem. 1992, 17, 474−480. 136, 15809−15812.
(21) Berke, H.; Hoffmann, R. Organometallic migration reactions. J. (42) Farrell, J. M.; Posaratnanathan, R. T.; Stephan, D. W. A family
Am. Chem. Soc. 1978, 100, 7224−7236. of N-heterocyclic carbene-stabilized borenium ions for metal-free
(22) West, N. M.; Miller, A. J. M.; Labinger, J. A.; Bercaw, J. E. imine hydrogenation catalysis. Chem. Sci. 2015, 6, 2010−2015.
Homogeneous syngas conversion. Coord. Chem. Rev. 2011, 255, 881− (43) Sitte, N. A.; Bursch, M.; Grimme, S.; Paradies, J. Frustrated
898. Lewis Pair Catalyzed Hydrogenation of Amides: Halides as Active
(23) Mahajan, D. Atom-economical reduction of carbon monoxide Lewis Base in the Metal-Free Hydrogen Activation. J. Am. Chem. Soc.
to methanol catalyzed by soluble transition metal complexes at low 2019, 141, 159−162.
temperatures. Top. Catal. 2005, 32, 209−214. (44) Sapsford, J. S.; Csókás, D.; Turnell-Ritson, R. C.; Parkin, L. A.;
(24) Zell, T.; Ben-David, Y.; Milstein, D. Unprecedented iron- Crawford, A. D.; Pápai, I.; Ashley, A. E. Transition Metal-Free Direct
catalyzed ester hydrogenation. Mild, selective, and efficient hydro- Hydrogenation of Esters via a Frustrated Lewis Pair. ACS Catal. 2021,
genation of trifluoroacetic esters to alcohols catalyzed by an iron 11, 9143−9150.
pincer complex. Angew. Chem., Int. Ed. 2014, 53, 4685−4689. (45) Guo, X.; Unglaube, F.; Kragl, U.; Mejia, E. B(C6F5)3-Catalyzed
(25) Srimani, D.; Mukherjee, A.; Goldberg, A. F.; Leitus, G.; Diskin- transfer hydrogenation of esters and organic carbonates towards
Posner, Y.; Shimon, L. J.; Ben David, Y.; Milstein, D. Cobalt-catalyzed alcohols with ammonia borane. Chem. Commun. 2022, 58, 6144−
hydrogenation of esters to alcohols: unexpected reactivity trend 6147.
indicates ester enolate intermediacy. Angew. Chem., Int. Ed. 2015, 54, (46) Li, Z.; Wang, Z.; Chekshin, N.; Qian, S.; Qiao, J. X.; Cheng, P.
12357−12360.
T.; Yeung, K. S.; Ewing, W. R.; Yu, J. Q. A tautomeric ligand enables
(26) Das, U. K.; Kumar, A.; Ben-David, Y.; Iron, M. A.; Milstein, D.
directed C-H hydroxylation with molecular oxygen. Science 2021, 372,
Manganese Catalyzed Hydrogenation of Carbamates and Urea
1452−1457.
Derivatives. J. Am. Chem. Soc. 2019, 141, 12962−12966.
(47) Rumo, C.; Stein, A.; Klehr, J.; Tachibana, R.; Prescimone, A.;
(27) Cabrero-Antonino, J. R.; Adam, R.; Papa, V.; Beller, M.
Häussinger, D.; Ward, T. R. An Artificial Metalloenzyme Based on a
Homogeneous and heterogeneous catalytic reduction of amides and
Copper Heteroscorpionate Enables sp3 C-H Functionalization via
related compounds using molecular hydrogen. Nat. Commun. 2020,
11, 3893. Intramolecular Carbene Insertion. J. Am. Chem. Soc. 2022, 144,
(28) Gausas, L.; Kristensen, S. K.; Sun, H.; Ahrens, A.; Donslund, B. 11676−11684.
S.; Lindhardt, A. T.; Skrydstrup, T. Catalytic Hydrogenation of (48) Lentz, N.; Reuge, S.; Albrecht, M. NADH-Type Hydride
Polyurethanes to Base Chemicals: From Model Systems to Storage and Release on a Functional Ligand for Efficient and Selective
Commercial and End-of-Life Polyurethane Materials. JACS Au Hydrogenation Catalysis. ACS Catal. 2023, 13, 9839−9844.
2021, 1, 517−524. (49) Chen, M.-W.; Wu, B.; Liu, Z.; Zhou, Y.-G. Biomimetic
(29) Liu, X.; Werner, T. Indirect reduction of CO2 and recycling of Asymmetric Reduction Based on the Regenerable Coenzyme
polymers by manganese-catalyzed transfer hydrogenation of amides, NAD(P)H Models. Acc. Chem. Res. 2023, 56, 2096−2109.
carbamates, urea derivatives, and polyurethanes. Chem. Sci. 2021, 12, (50) Podolyan, Y.; Gorb, L.; Leszczynski, J. Protonation of Nucleic
10590−10597. Acid Bases. A Comprehensive Post-Hartree−Fock Study of the
(30) Ryabchuk, P.; Stier, K.; Junge, K.; Checinski, M. P.; Beller, M. Energetics and Proton Affinities. J. Phys. Chem. A 2000, 104, 7346−
Molecularly Defined Manganese Catalyst for Low-Temperature 7352.
Hydrogenation of Carbon Monoxide to Methanol. J. Am. Chem. (51) Sanchez, R.; Giuliano, B. M.; Melandri, S.; Favero, L. B.;
Soc. 2019, 141, 16923−16929. Caminati, W. Gas-Phase Tautomeric Equilibrium of 4-Hydroxypyr-
(31) Kaithal, A.; Werlé, C.; Leitner, W. Alcohol-Assisted Hydro- imidine with Its Ketonic Forms: A Free Jet Millimeterwave
genation of Carbon Monoxide to Methanol Using Molecular Spectroscopy Study. J. Am. Chem. Soc. 2007, 129, 6287−6290.
Manganese Catalysts. JACS Au 2021, 1, 130−136. (52) Caminati, W. Nucleic acid bases in the gas phase. Angew. Chem.,
(32) Kumar, A.; Daw, P.; Milstein, D. Homogeneous Catalysis for Int. Ed. 2009, 48, 9030−9033.
Sustainable Energy: Hydrogen and Methanol Economies, Fuels from (53) Jin, L.; Lv, M.; Zhao, M.; Wang, R.; Zhao, C.; Lu, J.; Wang, L.;
Biomass, and Related Topics. Chem. Rev. 2022, 122, 385−441. Wang, W.; Wei, Y. Formic acid catalyzed isomerization of protonated
(33) Weetman, C.; Inoue, S. The Road Travelled: After Main-Group cytosine: a lower barrier reaction for tautomer production of potential
Elements as Transition Metals. ChemCatChem. 2018, 10, 4213−4228. biological importance. Phys. Chem. Chem. Phys. 2017, 19, 13515−
(34) Paradies, J. Structure-Reactivity Relationships in Borane-Based 13523.
FLP-Catalyzed Hydrogenations, Dehydrogenations, and Cycloisome- (54) Kasende, O. E.; Muya, J. T.; Broeckaert, L.; Maes, G.;
rizations. Acc. Chem. Res. 2023, 56, 821−834. Geerlings, P. Theoretical Study of the Regioselectivity of the
(35) Fujimori, S.; Inoue, S. Carbon Monoxide in Main-Group Interaction of 3-Methyl-4-pyrimidone and 1-Methyl-2-pyrimidone
Chemistry. J. Am. Chem. Soc. 2022, 144, 2034−2050. with Lewis Acids. J. Phys. Chem. A 2012, 116, 8608−8614.
(36) Jupp, A. R.; Stephan, D. W. New Directions for Frustrated (55) Houk, K. N.; Cheong, P. H. Computational prediction of small-
Lewis Pair Chemistry. Trends Chem. 2019, 1, 35−48. molecule catalysts. Nature 2008, 455, 309−313.
(37) Stephan, D. W. Diverse Uses of the Reaction of Frustrated (56) Guan, Y.; Wheeler, S. E. Automated Quantum Mechanical
Lewis Pair (FLP) with Hydrogen. J. Am. Chem. Soc. 2021, 143, Predictions of Enantioselectivity in a Rhodium-Catalyzed Asymmetric
20002−20014. Hydrogenation. Angew. Chem., Int. Ed. 2017, 56, 9101−9105.

I https://doi.org/10.1021/acsorginorgau.3c00064
ACS Org. Inorg. Au XXXX, XXX, XXX−XXX
ACS Organic & Inorganic Au pubs.acs.org/orginorgau Article

(57) Jiang, B.; Zhang, Q.; Dang, L. Theoretical studies on bridged (73) Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. Self-
frustrated Lewis pair (FLP) mediated H2 activation and CO2 consistent molecular orbital methods. XX. A basis set for correlated
hydrogenation. Org. Chem. Front. 2018, 5, 1905−1915. wave functions. J. Chern. Phys. 1980, 72, 650−654.
(58) Kondo, M.; Tatewaki, H.; Masaoka, S. Design of molecular (74) Tomasi, J.; Persico, M. Molecular Interactions in Solution: An
water oxidation catalysts with earth-abundant metal ions. Chem. Soc. Overview of Methods Based on Continuous Distributions of the
Rev. 2021, 50, 6790−6831. Solvent. Chem. Rev. 1994, 94, 2027−2094.
(59) Mroz, A. M.; Posligua, V.; Tarzia, A.; Wolpert, E. H.; Jelfs, K. E. (75) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal
Into the Unknown: How Computation Can Help Explore Uncharted Solvation Model Based on Solute Electron Density and on a
Material Space. J. Am. Chem. Soc. 2022, 144, 18730−18743. Continuum Model of the Solvent Defined by the Bulk Dielectric
(60) Tu, Z.; Stuyver, T.; Coley, C. W. Predictive chemistry: machine Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113,
learning for reaction deployment, reaction development, and reaction 6378−6396.
discovery. Chem. Sci. 2023, 14, 226−244. (76) Tsutsumi, T.; Ono, Y.; Arai, Z.; Taketsugu, T. Visualization of
(61) Li, L.; Lei, M.; Xie, Y.; Schaefer, H. F., III; Chen, B.; Hoffmann, the Intrinsic Reaction Coordinate and Global Reaction Route Map by
R. Stabilizing a different cyclooctatetraene stereoisomer. Proc. Natl. Classical Multidimensional Scaling. J. Chem. Theory Comput. 2018, 14,
Acad. Sci. U.S.A. 2017, 114, 9803−9808. 4263−4270.
(62) Li, L.; Wu, Z.; Zhu, H.; Robinson, G. H.; Xie, Y.; Schaefer, H. (77) Glendening, E. D.; Landis, C. R.; Weinhold, F. NBO 7.0: New
F. Reduction of Dinitrogen via 2,3′-Bipyridine-Mediated Tetrabora- vistas in localized and delocalized chemical bonding theory. J. Comput.
tion. J. Am. Chem. Soc. 2020, 142, 6244−6250. Chem. 2019, 40, 2234−2241.
(63) Zahrt, A. F.; Henle, J. J.; Rose, B. T.; Wang, Y.; Darrow, W. T.; (78) Chen, Z.; Wannere, C. S.; Corminboeuf, C.; Puchta, R.;
Denmark, S. E. Prediction of higher-selectivity catalysts by computer- Schleyer, P. V. R. Nucleus-Independent Chemical Shifts (NICS) as an
driven workflow and machine learning. Science 2019, 363, Aromaticity Criterion. Chem. Rev. 2005, 105, 3842−3888.
(79) Qin, J.; Li, F.; Qiu, R.; Chen, L.; Luo, L.; Wang, M.; Pu, Z.;
No. eaau5631.
Shuai, M. Insights into the Metal-CO Bond in O2M(η1-CO) (M = Cr,
(64) Hannagan, R. T.; Giannakakis, G.; Réocreux, R.; Schumann, J.;
Mo, W, Nd, and U) Complexes. Inorg. Chem. 2022, 61, 2066−2075.
Finzel, J.; Wang, Y.; Michaelides, A.; Deshlahra, P.; Christopher, P.;
(80) Zhang, S.-Q.; Taylor, B. L. H.; Ji, C.-L.; Gao, Y.; Harris, M. R.;
Flytzani-Stephanopoulos, M.; Stamatakis, M.; Sykes, E. C. H. First-
Hanna, L. E.; Jarvo, E. R.; Houk, K. N.; Hong, X. Mechanism and
principles design of a single-atom−alloy propane dehydrogenation Origins of Ligand-Controlled Stereoselectivity of Ni-Catalyzed
catalyst. Science 2021, 372, 1444−1447. Suzuki-Miyaura Coupling with Benzylic Esters: A Computational
(65) Sanchez, D. M.; Raucci, U.; Martínez, T. J. In Silico Discovery Study. J. Am. Chem. Soc. 2017, 139, 12994−13005.
of Multistep Chemistry Initiated by a Conical Intersection: The (81) Chen, Y.; Zhang, X.; Liu, F.; He, G.; Zhang, J.; Houk, K. N.;
Challenging Case of Donor-Acceptor Stenhouse Adducts. J. Am. Smith, A. B., III; Liang, Y. The role of CuI in the siloxane-mediated
Chem. Soc. 2021, 143, 20015−20021. Pd-catalyzed cross-coupling reactions of aryl iodides with aryl lithium
(66) Kim, S.-T.; Strauss, M. J.; Cabré, A.; Buchwald, S. L. Room- reagents. Chin. Chem. Lett. 2021, 32, 441−444.
Temperature Cu-Catalyzed Amination of Aryl Bromides Enabled by (82) Qin, C.; Koengeter, T.; Zhao, F.; Mu, Y.; Liu, F.; Houk, K. N.;
DFT-Guided Ligand Design. J. Am. Chem. Soc. 2023, 145, 6966− Hoveyda, A. H. Z-Trisubstituted α,β-Unsaturated Esters and Acid
6975. Fluorides through Stereocontrolled Catalytic Cross-Metathesis. J. Am.
(67) Xu, R.; Meisner, J.; Chang, A. M.; Thompson, K. C.; Martinez, Chem. Soc. 2023, 145, 3748−3762.
T. J. First principles reaction discovery: from the Schrodinger (83) Kaasik, M.; Chen, P.-P.; Ričko, S.; Jo̷ rgensen, K. A.; Houk, K.
equation to experimental prediction for methane pyrolysis. Chem. Sci. N. Asymmetric [4 + 2], [6 + 2], and [6 + 4] Cycloadditions of
2023, 14, 7447−7464. Isomeric Formyl Cycloheptatrienes Catalyzed by a Chiral Diamine
(68) Jia, Z.; Li, L.; Zhang, X.; Yang, K.; Li, H.; Xie, Y.; Schaefer, H. Catalyst. J. Am. Chem. Soc. 2023, 145, 23874−23890.
F., III Acceleration Effect of Bases on Mn Pincer Complex-Catalyzed (84) Galeotti, M.; Lee, W.; Sisti, S.; Casciotti, M.; Salamone, M.;
CO2 Hydroboration. Inorg. Chem. 2022, 61, 3970−3980. Houk, K. N.; Bietti, M. Radical and Cationic Pathways in C(sp3)-H
(69) Rong, H.; Zhang, Y.; Ai, X.; Li, W.; Cao, F.; Li, L. Theoretical Bond Oxygenation by Dioxiranes of Bicyclic and Spirocyclic
Study on the Hydrogenolysis of Polyurethanes to Improve the Hydrocarbons Bearing Cyclopropane Moieties. J. Am. Chem. Soc.
Catalytic Activities. Inorg. Chem. 2022, 61, 14662−14672. 2023, 145, 24021−24034.
(70) Ai, X.; Xie, X.; Song, X.; Li, L.; Schaefer, H. F., III Is the (85) Ryu, H.; Park, J.; Kim, H. K.; Park, J. Y.; Kim, S.-T.; Baik, M.-H.
polarization of the C = C bond imperative for bifunctional outer- Pitfalls in Computational Modeling of Chemical Reactions and How
sphere C = C hydrogenation? Org. Chem. Front. 2023, 10, 1301− To Avoid Them. Organometallics 2018, 37, 3228−3239.
1308.
(71) Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Hybrid
Density Functionals with Damped Atom-Atom Dispersion Correc-
tions. Phys. Chem. Chem. Phys. 2008, 10, 6615−6620.
(72) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci,
B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H.
P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.;
Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.;
Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.;
Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.;
Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi,
R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar,
S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox,
J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.;
Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.;
Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.;
Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A.
D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J.
Gaussian 09, revision B.01; Gaussian, Inc.: Wallingford, CT, 2009.

J https://doi.org/10.1021/acsorginorgau.3c00064
ACS Org. Inorg. Au XXXX, XXX, XXX−XXX

You might also like