Articulo 3-Catalisis Heterogenea

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Molecular Catalysis 467 (2019) 1–8

Contents lists available at ScienceDirect

Molecular Catalysis
journal homepage: www.elsevier.com/locate/mcat

Molecular mechanism of heterogeneous supramolecular catalysis of metal- T


free cucurbituril solid for epoxide alcoholysis
Lina Xua, Guoyong Fanga,b, , Yanghong Yua, Yuefan Maa, Zihang Yea, Zhenyu Lib,
⁎ ⁎⁎

a
Key Laboratory of Carbon Materials of Zhejiang Province, College of Chemistry and Materials Engineering, Wenzhou University, Wenzhou 325035, China
b
Hefei National Laboratory of Physical Sciences at the Microscale, School of Chemistry and Materials Sciences, University of Science and Technology of China, Hefei
230026, China

ARTICLE INFO ABSTRACT

Keywords: Heterogeneous and supramolecular catalysis are fundamental processes in chemistry. To understand the sy-
Supramolecular catalysis nergistic role between heterogeneous and supramolecular catalysis, the catalytic mechanism of cucurbituril solid
Heterogeneous catalysis for epoxide alcoholysis was investigated by performing density functional theory (DFT) calculations. The results
Cucurbituril reveal that styrene oxide (StyOx) ethanolysis has an inherent regioselectivity, which results from different
Epoxide alcoholysis
groups linked to the epoxy group. The hydronium ion can catalyze the ring-opening of StyOx ethanolysis and
Density functional theory calculation
lead to the formation of a planar carbenium ion. StyOx can also be hydrolyzed via the homogenous catalysis of
acid to produce 1, 2-diol. Cucurbituril solid can catalyze epoxide alcoholysis because of its acidic property. Its
unique cavity can lead to a favorable frontside-attack of the alcohol on the carbenium ion. The product from the
heterogeneous catalysis of cucurbituril solid is pure β-alkoxy alcohol. The results are important to understand
heterogeneous and supramolecular catalysis and the design of new and effective supramolecular catalysts.

1. Introduction recognize suitable guest molecules and catalyze many chemical reac-
tions. CBn have been used to control and catalyze various chemical
In the past decade, extensive progress has been made in supramo- reactions in the liquid phase, including photochemical and thermal
lecular catalysis, which has become a hot topic in the field of chemistry reactions, such as photodimerization, photolysis, cycloaddition, hy-
[1–12]. In principle, supramolecular catalysis is similar to enzyme drolysis, oxidation, and desilylation [34–50]. Recently, the catalytic
catalysis and provides the cavities and catalytic sites for various reac- properties of CBn in the gas and solid phases have been investigated
tions [13–19]. Because of limitations of experimental and computa- [24,51,52].
tional conditions, the molecular mechanism of supramolecular catalysis To obtain insight into the catalytic mechanism, a few theoretical
is still unclear. To date, only a few supramolecule-catalyzed reactions calculations have been performed to illustrate the nature of cucurbi-
exist, such as imine formation, methylation and cycloaddition, which turil-based supramolecular catalysis [24–26]. The retro-Diels–Alder
have been studied in detail by using accurate quantum-chemistry cal- reaction inside cucurbiturils in the gas phase has been explored theo-
culations [20–26]. retically [24]. The enhanced reactivity of 1, 3-dipolar cycloaddition
As a well-known supramolecular container, pumpkin-shaped cu- between an azide and an alkyne inside cucurbiturils in the liquid phase
curbiturils are a family of macrocyclic compounds that comprise gly- via homogeneous catalysis has also been addressed [25,26]. However, a
coluril units that are connected by methylene (eCH2e) bridges limited understanding exists of the supramolecular catalytic mechanism
[27–33]. Cucurbiturils are written commonly as cucurbit[n]urils, where and nature of cucurbituril solids that are produced by heterogeneous
n is the number of glycoluril units (]C4H2N4O2]), and they are ab- catalysis.
breviated as CBn or CB[n] (n = 5 ∼ 10). They can be synthesized by Epoxides, which are a type of heterocyclic compound, are used
the condensation of glycoluril and formaldehyde under acidic condi- extensively in organic compounds and as important intermediates in
tions (Scheme 1). Because of their unique cavity, cucurbiturils can pharmaceutical synthesis [53–55]. Epoxide ring-opening with alcohols


Corresponding author at: Key Laboratory of Carbon Materials of Zhejiang Province, College of Chemistry and Materials Engineering, Wenzhou University,
Wenzhou 325035, China.
⁎⁎
Corresponding author at: Hefei National Laboratory of Physical Sciences at the Microscale, School of Chemistry and Materials Sciences, University of Science and
Technology of China, Hefei 230026, China.
E-mail addresses: fanggy@wzu.edu.cn (G. Fang), zyli@ustc.edu.cn (Z. Li).

https://doi.org/10.1016/j.mcat.2019.01.021
Received 11 December 2018; Received in revised form 11 January 2019; Accepted 21 January 2019
Available online 25 January 2019
2468-8231/ © 2019 Elsevier B.V. All rights reserved.
L. Xu et al. Molecular Catalysis 467 (2019) 1–8

P2, nocat, and H3O+, represent different pathways and catalytic con-
ditions.
Fig. 1 shows the Gibbs free-energy profiles of two reaction pathways
of StyOx and ethanol without a catalyst. The corresponding products
are 2-ethoxy-2-phenylethanol (P1, StyOEt) and 2-ethoxy-1-pheny-
lethanol (P2), respectively. The former is a useful β-alkoxy alcohol. The
difference between the two pathways is that the hydroxyl group of
ethanol can attack different CeO bonds (C1eO3 and C2eO3, shown in
Scheme 2) of the epoxy group of styrene oxide, which results in StyOx
regioselectivity. They can form a concerted four-membered ring (4MR)
Scheme 1. Synthesis and chemical structure of cucurbit[n]urils. transition state (TS) that is composed of epoxy and hydroxyl groups.
Finally, they can be converted into β-ethoxy alcohol products. During
the ring-opening of the first P1 pathway, the old CeO bond of the epoxy
is an important synthesis technique for β-alkoxy alcohols, which are group is lengthened from 1.43 Å (Im1) to 2.23 Å (TS1) and 2.44 Å (P1),
useful intermediates and can be used to generate α-alkoxy ketones, α- and that of the old OeH bond of the hydroxyl group is lengthened from
alkoxy acids, and many natural products and drugs. We investigated the 0.97 Å (Im1) to 1.03 Å (TS1) and 2.35 Å (P1). The distance of the C and
catalytic mechanism of metal-free cucurbituril solids for epoxide alco- O atoms of the reaction center is reduced from 3.11 Å (Im1) to 2.64 Å
holysis. It is expected that insights into the catalytic mechanism of (TS1) and 1.42 Å (P1). That of the O and H atoms is shortened from
cucurbituril solid may increase the knowledge of heterogeneous and 2.02 Å (Im1) to 1.50 Å (TS1) and 0.96 Å (P1). These results indicate
supramolecular catalysis. that the epoxy group is opened, the hydroxyl group is broken, and new
CeO and OeH bonds are formed. The second P2 pathway is similar to
2. Computational details the first. From an energy perspective, the two products, P1 and P2, are
almost similarly exergonic by 15.3 and 15.0 kcal/mol, respectively.
All stationary points for the alcoholysis reactions of StyOx without a However, the two regioselective pathways require high activation free
catalyst and with an acidic catalyst and cucurbituril solid were opti- energies of 45.0 and 59.3 kcal/mol, which indicates that the two
mized fully within the framework of density functional theory (DFT) in pathways cannot occur in the absence of a catalyst.
the Gaussian 09 program [56]. The 6–311 G(d,p) basis set and GD3
dispersion-corrected M06-2X functional (M06-2X-GD3) were used for 3.2. Alcoholysis of StyOx with acid catalyst
all calculations, which can describe the noncovalent interaction
[57–61]. All geometry optimizations were performed in ethanol solvent It is difficult to react StyOx and ethanol without a catalyst. Aqueous
with a solvation model based on density [62,63]. This solution-phase acid can be used as a catalyst for the ring-opening of epoxides [65,66].
optimization method can yield a reasonable free-energy surface of so- The reaction mechanism of acid-catalyzed ethanolysis of StyOx is
lution reactions [62]. To simulate the condition for heterogeneous considered further.
catalysis, all atoms of cucurbituril solid were fixed in geometry opti- Fig. 2 show a possible stepwise mechanism for the reaction of StyOx
mizations. The Gibbs free energies of all species were calculated at and ethanol with an acid (hydronium ion, H3O+) catalyst. H3O+ can
298.15 K and 1 atm and corrected by 1.9 kcal/mol to convert 1 atm to interact with StyOx and ethanol to form a hydrogen-bonding inter-
1 M standard state [63]. The vibrational frequency and intrinsic reac- mediate, Im1P1,H3O+, in which the epoxy group is almost protonated
tion coordinate (IRC) calculations were carried out to verify whether a by H3O+ with a distance of 1.14 Å, which facilitates the ring-opening of
minimum or a transition state exists [64]. StyOx. It can undergo a transition state of the ring-opening of the epoxy
group, TS1P1,H3O+, in which the proton is transferred entirely to the
3. Results and discussion epoxy group and the C1-O3 bond is lengthened from 1.47 Å to 1.98 Å.
The reaction can then proceed via an intermediate, Im2P1,H3O+, which
3.1. Alcoholysis of epoxides without catalyst comprises water, ethanol, and carbenium ion, in which the carbenium
ion is planar and is located 3.54 Å from the ethanol hydroxyl group.
In general, the ring-opening reactions of epoxide alcoholysis require Afterwards, the ethanol molecule can attack the planar carbenium ion
catalysts, such as acids, bases, or zeolites [65,66]. For example, etha- via a low-energy transition state, TS2P1,H3O+, and the distance of the C
nolysis of StyOx occurs with difficulty without a catalyst, and includes atom of the carbenium ion and the O atom of ethanol can be reduced to
two possible regioselective pathways (P1 and P2) (Scheme 2). Herein, 3.26 Å. Lastly, a new CeO bond and hydrogen-bonding intermediate,
alcoholysis of StyOx is used only as a reference reaction and facilitates Im3P1,H3O+, is formed and the product, 2-ethoxy-2-phenylethanol
the discussion of subsequent mechanisms with acid and cucurbituril (P1), is deprotonated. The H3O+ is recycled.
catalysts. To represent the stationary points along all the pathways in a Because of its planar property, the carbenium ion is susceptible to
concise way, we use denotations of intermediates Im1 and Im2, tran- frontside and backside-attacks by ethanol. Two products (P1 and P1’)
sition states TS1 and TS2, and products P1 and P2. Superscripts, P1 and of 2-ethoxy-2-phenylethanol are enantiomers and mirror images of each

Scheme 2. Two regioselective pathways (P1 and P2) of StyOx ethanolysis.

2
L. Xu et al. Molecular Catalysis 467 (2019) 1–8

Fig. 1. Two options for the regioselective ring-opening ethanolysis of StyOx. Values represent the distances between two atoms (Å).

other. This property is termed epoxide stereoselectivity. In homo- occur easily.


geneous catalysis, ethanol exists in a free state and can attack the car- Fig. 3 shows another regioselective pathway for the reaction of
benium ions from the front and the back. So, the stereoselectivity of StyOx and ethanol with a H3O+ catalyst. This pathway is similar to the
StyOx alcoholysis is difficult to control. P2 pathway without a catalyst and the open C2eO3 bond of the epoxy
The rate-determining step of the entire pathway is a H3O+-cata- group. Because the C2eO3 bond is not linked to the conjugated benzyl
lyzed ring-opening of StyOx with an energy barrier of 9.4 kcal/mol. group, this pathway does not undergo an intermediate of the planar
Ethanol attack of the planar carbenium ion is almost barrierless and carbenium ion. First, H3O+, StyOx, and ethanol can form a strong hy-
requires only 1.4 kcal/mol. The energy barrier for product release is drogen-bonding intermediate, Im1P2,H3O+, in which the ethanol hy-
only 8.5 kcal/mol. Therefore, H3O+-catalyzed epoxide alcoholysis can droxyl group is located on one side of the C2eO3 bond of the epoxy

Fig. 2. Two stereoselective H3O+-catalyzed pathways (H3O+-catalyzed P1 and H3O+-catalyzed P1’) of StyOx ethanolysis via the first regioselective pathway. Values
represent the distances between two atoms (Å).

3
L. Xu et al. Molecular Catalysis 467 (2019) 1–8

energy barriers of TS1P1,H3O+ that result in β-alkoxy alcohol (P1 + P1')


and TS1P3,H3O+ that result in 1, 2-diol (P3 + P3') are 7.4 and 5.2 kcal/
mol, respectively. According to the Arrhenius equation [68], this
2.2 kcal/mol energy difference indicates that the reaction under con-
sideration gives β-alkoxy alcohol as the major product with a trace
amount of 1, 2-diol.

3.3. Ethanolysis of StyOx in CB7 solid phase

As well-known supramolecular catalysts, CBn has been used in


heterogeneous catalysis [51,52]. In general, cucurbiturils are synthe-
sized under acidic reaction conditions [28–33]. The crystal structures of
cucurbiturils show that H3O+ is located at the negatively charged
carbonyl portals and the corresponding anion lies at the CBn voids
[69–72]. Based on the shape and size of the host and guest molecules,
we selected one cucurbituril, cucurbit[7]uril (CB7), and investigated
the ethanolysis mechanism of StyOx in CB7 solid. As shown in Fig. 5,
H3O+ and CB7 can form a protonated CB7 (CB7·H3O+) via two strong
hydrogen bonds with a distance of 1.50 Å and 1.51 Å, respectively.
Because of its unique cavity and acidic property, CBn can be used as a
solid acid catalyst to catalyze the ethanolysis of StyOx.
Fig. 3. Another regioselective H3O+-catalyzed pathway (H3O+-catalyzed P2) Fig. 5 shows the Gibbs free energy profile of frontside-attack etha-
of StyOx ethanolysis. Values represent the distances between two atoms (Å).
nolysis of StyOx in protonated CB7. The catalytic mechanism is similar
to the acid-catalyzed mechanism. Two guest molecules, ethanol and
group. This group undergoes a concerted transition state, TS1P2,H3O+, StyOx, enter the host molecule and protonate CB7, which can lead to
in which the epoxy group is opened and a new CeO bond is formed. the formation of a stable intermediate, Im1P1,CB7·H3O+. In
The old OeH bond is broken and a new OeH bond is formed via proton Im1P1,CB7·H3O+, the epoxy group of StyOx and H3O+ at the carbonyl
transfer. Finally, a hydrogen-bonding intermediate, Im2P2,H3O+, forms portal of CB7 can form a strong 1.47-Å hydrogen bond, which facil-
and the product, 2-ethoxy-1-phenylethanol (P2), is deprotonated. itates subsequent ring-opening. The ethanol molecule can locate im-
The energy barrier of the H3O+-catalyzed P2 pathway can be mediately below H3O+ and the epoxy group. Subsequently, the epoxy
˜30.0 kcal/mol lower than that of the P2 pathway without a catalyst group of the StyOx in the protonated CB7 can be opened because of
and decreased from 59.3 kcal/mol to 29.5 kcal/mol, which indicates H3O+ via a transition state, TS1P1,CB7·H3O+, where the epoxy group is
that H3O+ can accelerate the chemical reaction. Compared with the protonated completely and the distance of the CeO bond of the epoxy
H3O+-catalyzed P1 pathway, the H3O+-catalyzed P2 pathway does not group increases from 1.44 Å to 1.94 Å. This can result in the formation
produce a planar carbenium ion intermediate and is unfavorable. These of an intermediate, Im2P1,CB7·H3O+, between the carbenium ion and
results show that the regioselectivity for β-alkoxy alcohols is an in- ethanol in the CB7 cavity. The ethanol molecule can attack the carbe-
herent character of StyOx ethanolysis and stems from different groups nium ion from the front via a mostly barrierless transition state,
that are linked to the epoxy group. This finding is also consistent with TS2P1,CB7·H3O+, in which the distance of the C atom of the carbenium
the experimental results, where the product is the unique β-alkoxy al- ion and the O atom of ethanol decreases from 3.16 Å to 2.24 Å. Lastly,
cohol [52]. In the ensuing studies, only the first regioselective pathway the intermediate, Im3P1,CB7·H3O+, and β-ethoxy alcohol (P1) in the CB7
(P1) is considered. cavity is formed and the product is released from the CB7 cavity. The
The H3O+-catalyzed ethanolysis of StyOx is a hypothetical situa- protonated CB7 (CB7·H3O+) is recycled.
tion. If H3O+ is present in the ethanolysis, then water (H2O) can par- In the frontside-attack pathway in CB7, the barriers of the ring-
ticipate in the chemical reaction. As a nucleophile, H2O can attack the opening of the epoxy group of StyOx and the attack of ethanol on the
carbenium ion and compete with the ethanol. The product is a diol, 1- planar carbenium ion are 11.3 and 1.2 kcal/mol, respectively. Because
phenylethane-1, 2-diol (P3, Scheme 3). The reaction is the hydrolysis of of the strong interaction between CB7 and the guest molecules, the free
StyOx and the product has no regioselectivity, which differs from the energies of all the intermediates and transition states are lower than
alcoholysis of StyOx. that of the reactants. The rate-determining step of the CB7-catalyzed
Fig. 4 shows the H3O+-catalyzed hydrolysis mechanism of StyOx, pathway is the release of β-ethoxy alcohol from the solid CB7 and the
which is similar to the catalyzed ethanolysis mechanism that includes barrier is 15.8 kcal/mol. These results indicate that the unique cavity of
the ring-opening of the epoxy group, the attack of the hydroxyl group CBn and H3O+ at the portal of the solid CB7 can catalyze the epoxide
on the carbenium ion, and the release of the product. Because of the alcoholysis synergistically.
planar property and stereoselectivity of the carbenium ion, the hydro- In the cavity of the solid CB7, ethanol is also located on another side
lyzed product of StyOx, 1-phenylethane-1, 2-diol, also includes two of H3O+ and the epoxy group of StyOx. Fig. 6 shows the backside-
enantiomers (P3 and P3’). As shown by the forward energy barriers (7.9 attack ethanolysis of StyOx in protonated CB7. Frontside or backside-
and 8.1 kcal/mol) of the ring-opening of the epoxy group and the attack attack results in H3O+ at the portal of CB7 interacting strongly with the
of nucleophiles on the carbenium ion in Figs. 2 and 4, the ethanolysis of epoxy group, which can decrease the energy barrier and catalyze the
StyOx and H3O+-catalyzed hydrolysis can compete. However, on the ring-opening of StyOx significantly. For frontside-attack, the ethanol
basis of the principle of microscopic reversibility [67], the reverse molecule and the epoxy group of StyOx are located on one side, as
shown in Im1P1,CB7·H3O+ of Fig. 5. For backside-attack, the ethanol

Scheme 3. Hydrolysis reaction pathway of StyOx.

4
L. Xu et al. Molecular Catalysis 467 (2019) 1–8

Fig. 4. Two regioselective H3O+-cata-


lyzed pathways (H3O+-catalyzed P3
and H3O+-catalyzed P3’) of StyOx hy-
drolysis. Values represent the distances
between two atoms (Å).

Fig. 5. Frontside-attack ethanolysis of StyOx in solid CB7. Values represent the distances between two atoms (Å).

molecule and the epoxy group of StyOx are located on two sides, as product P1’ of the backside-attack pathway is an enantiomer of the
shown in Im1P1’,CB7·H3O+ of Fig. 6. frontside-attack pathway. From Im3P1’,CB7·H3O+, the energy barrier of
Similar to the frontside-attack pathway, backside-attack ethanolysis the release of the product is up to 24.6 kcal/mol. Furthermore, the
also includes the ring-opening of StyOx, the attack of ethanol on the backside-attack pathway is endergonic by 4.2 kcal/mol, which is
planar carbenium ion, and product release. The entire backside-attack 19.5 kcal/mol higher than that in the frontside-attack pathway. These
pathway passes through Im1P1’,CB7·H3O+, TS1P1’,CB7·H3O+, results indicate that the backside-attack pathway is thermodynamically
Im2P1’,CB7·H3O+, TS2P1’,CB7·H3O+, and Im3P1’,CB7·H3O+. However, the and kinetically unfavorable. The reason for this behavior is that the

5
L. Xu et al. Molecular Catalysis 467 (2019) 1–8

Fig. 6. Backside-attack ethanolysis of StyOx in solid CB7. Values represent the distances between two atoms (Å).

release of the product damages the structure of the initial protonated


CB7 and the catalyst cannot be recovered to the initial CB7·H3O+ state.
In addition, as shown in Fig. 6, after the ring-opening of StyOx, the
H2O molecule, which was converted from H3O+ and had absorbed at
the portal of CB7, may attack the planar carbenium ion, which results
in the formation of 1-phenylethane-1, 2-diol (P3). Subsequently, the
H2O-attacked pathway produces Im2P3,CB7·H3O+, TS2P3,CB7·H3O+, and
Im3P3,CB7·H3O+. The release of product P3 requires a high barrier of
31.9 kcal/mol. Furthermore, this H2O-attacked pathway is endergonic
by 3.7 kcal/mol, which is 19.0 kcal/mol higher than that in the front-
side-attack pathway. Product release in the H2O-attacked pathway also
damages the structure of the initial protonated CB7 and limits catalyst
regeneration. The H2O-attacked pathway in the solid CB7 is thermo-
dynamically and kinetically unfavorable. So, producing the diol (P3) in
the presence of CB7 is very difficult. This finding agrees with the het-
erogeneous experimental result, where the product of StyOx ethanolysis
catalyzed by CB7 solid is unique StyOEt and not the diol [52].
A comparison of Figs. 5 and 6 shows that the frontside-attack
pathway of StyOx ethanolysis in the solid CB7 is the most favorable Fig. 7. Probable catalytic cycle of StyOx ethanolysis by CB7 solid.
pathway (Fig. 7), which is different from the case via free H3O+ cata-
lysis mentioned above. In such a frontside-attack pathway, the epoxy formation of a planar carbenium ion. StyOx can be hydrolyzed via
group of StyOx can be regioselectively ring-opened by the protonated homogenous catalysis of acid to produce 1, 2-diol. The CB7 solid has an
CB7 via a strong hydrogen bonding and then stereoselectively attacked inherent acid property and can catalyze epoxide alcoholysis via het-
by EtOH. These results show that StyOx alcoholysis should have re- erogenous catalysis. The unique cavity of CB7 can lead to a favorable
gioselectivity and stereoselectivity via heterogeneous supramolecular frontside-attack of ethanol to StyOx. The product via heterogeneous
catalysis of the metal-free CB7 solid. Because the reactant, StyOx, has catalysis of CB7 solid is pure β-alkoxy alcohol. These findings on the
two enantiomers and CB7 has two symmetrical portals, it is difficult to selectivity and mechanism of heterogeneous and supramolecular cata-
control the stereoselectivity of StyOx ethanolysis. However, the product lysis of cucurbituril solid may help in the design of new and effective
must be regioselective and pure β-alkoxy alcohol [52]. supramolecular catalysts.

4. Conclusions Acknowledgements

In summary, the uncatalyzed, acid-catalyzed, and cucurbituril-cat- This work was supported by the National Natural Science
alyzed mechanisms of epoxide alcoholysis have been investigated by Foundation of China (21703158) and the Zhejiang Provincial Natural
performing DFT calculations. StyOx ethanolysis has an inherent re- Science Foundation of China (LQ15B030001). We thank Professors
gioselectivity, which stems from different groups linked to the epoxy Qing Xu and Jun Jiang of Wenzhou University for helpful discussions
group. H3O+ can catalyze the ring-opening of StyOx, and lead to the and the National Supercomputer Center in Guangzhou and the

6
L. Xu et al. Molecular Catalysis 467 (2019) 1–8

Supercomputing Center of University of Science and Technology of supramolecular catalysis within cucurbiturils, Chem. Eur. J. 18 (2012)
China for providing computing resources. 12178–12190.
[32] S.J. Barrow, S. Kasera, M.J. Rowland, J. del Barrio, O.A. Scherman, Cucurbituril-
based molecular recognition, Chem. Rev. 115 (2015) 12320–12406.
Appendix A. Supplementary data [33] K.I. Assaf, W.M. Nau, Cucurbiturils: from synthesis to high-affinity binding and
catalysis, Chem. Soc. Rev. 44 (2015) 394–418.
[34] S.Y. Jon, Y.H. Ko, S.H. Park, H.-J. Kim, K. Kim, A facile, stereoselective [2+2]
Supplementary material related to this article can be found, in the photoreaction mediated by cucurbit[8]uril, Chem. Commun. 37 (2001) 1938–1939.
online version, at doi:https://doi.org/10.1016/j.mcat.2019.01.021. [35] M. Pattabiraman, A. Natarajan, L.S. Kaanumalle, V. Ramamurthy, Templating
photodimerization of trans-cinnamic acids with cucurbit[8]uril and γ-cyclodextrin,
Org. Lett. 7 (2005) 529–532.
References [36] M. Pattabiraman, A. Natarajan, R. Kaliappan, J.T. Mague, V. Ramamurthy,
Template directed photodimerization of trans-1,2-bis(n-pyridyl)ethylenes and stil-
[1] R. Breslow, S.D. Dong, Biomimetic reactions catalyzed by cyclodextrins and their bazoles in water, Chem. Commun. 41 (2005) 4542–4544.
derivatives, Chem. Rev. 98 (1998) 1997–2011. [37] H. Yang, Z. Ma, Z.Q. Wang, X. Zhang, Fabricating covalently attached hyper-
[2] T.S. Koblenz, J. Wassenaar, J.N.H. Reek, Reactivity within a confined self-as- branched polymers by combining photochemistry with supramolecular poly-
sembled nanospace, Chem. Soc. Rev. 37 (2008) 247–262. merization, Polym. Chem. 5 (2014) 1471–1476.
[3] M.D. Pluth, R.G. Bergman, K.N. Raymond, Proton-mediated chemistry and catalysis [38] N. Barooah, B.C. Pemberton, J. Sivaguru, Manipulating photochemical reactivity of
in a self-assembled supramolecular host, Acc. Chem. Res. 42 (2009) 1650–1659. coumarins within cucurbituril nanocavities, Org. Lett. 10 (2008) 3339–3342.
[4] M. Yoshizawa, J.K. Klosterman, M. Fujita, Functional molecular flasks: new prop- [39] B.C. Pemberton, N. Barooah, D.K. Srivatsava, J. Sivaguru, Supramolecular photo-
erties and reactions within discrete, self-assembled hosts, Angew. Chem. Int. Ed. 48 catalysis by confinement-photodimerization of coumarins within cucurbit[8]urils,
(2009) 3418–3438. Chem. Commun. 46 (2010) 225–227.
[5] M.J. Wiester, P.A. Ulmann, C.A. Mirkin, Enzyme mimics based upon supramole- [40] B.C. Pemberton, R.K. Singh, A.C. Johnson, S. Jockusch, J.P.D. Silva, A. Ugrinov,
cular coordination chemistry, Angew. Chem. Int. Ed. 50 (2011) 114–137. N.J. Turro, D.K. Srivastava, J. Sivaguru, Supramolecular photocatalysis: insights
[6] H. Amouri, C. Desmarets, J. Moussa, Confined nanospaces in metallocages: guest into cucurbit[8]uril catalyzed photodimerization of 6-methylcoumarin, Chem.
molecules, weakly encapsulated anions, and catalyst sequestration, Chem. Rev. 112 Commun. 47 (2011) 6323–6325.
(2012) 2015–2041. [41] C. Yang, T. Mori, Y. Origane, Y.H. Ko, N. Selvapalam, K. Kim, Y. Inoue, Highly
[7] M. Raynal, P. Ballester, A. Vidal-Ferran, P.W.N.M. van Leeuwen, Supramolecular stereoselective photocyclodimerization of α-cyclodextrin-appended anthracene
catalysis. part 1: non-covalent interactions as a tool for building and modifying mediated by γ-cyclodextrin and cucurbit[8]uril: a dramatic steric effect operating
homogeneous catalysts, Chem. Soc. Rev. 43 (2014) 1660–1733. outside the binding site, J. Am. Chem. Soc. 130 (2008) 8574–8575.
[8] M. Raynal, P. Ballester, A. Vidal-Ferran, P.W.N.M. van Leeuwen, Supramolecular [42] A.L. Koner, C. Márquez, M.H. Dickman, W.M. Nau, Transition-metal-promoted
catalysis. Part 2: artificial enzyme mimics, Chem. Soc. Rev. 43 (2014) 1734–1787. chemoselective photoreactions at the cucurbituril rim, Angew. Chem. Int. Ed. 50
[9] C.J. Brown, F.D. Toste, R.G. Bergman, K.N. Raymond, Supramolecular catalysis in (2011) 545–548.
metal-ligand cluster hosts, Chem. Rev. 115 (2015) 3012–3035. [43] L. Zheng, S. Sonzini, M. Ambarwati, E. Rosta, O.A. Scherman, A. Herrmann, Turning
[10] S.H.A.M. Leenders, R. Gramage-Doria, B. de Bruin, J.N.H. Reek, Transition metal cucurbit[8]uril into a supramolecular nanoreactor for asymmetric catalysis, Angew.
catalysis in confined space, Chem. Soc. Rev. 44 (2015) 433–448. Chem. Int. Ed. 54 (2015) 13007–13011.
[11] A. Galan, P. Ballester, Stabilization of reactive species by supramolecular en- [44] W.L. Mock, T.A. Irra, J.P. Wepsiec, T.L. Manimaran, Cycloaddition induced by
capsulation, Chem. Soc. Rev. 45 (2016) 1720–1737. cucurbituril. a case of pauling principle catalysis, J. Org. Chem. 48 (1983)
[12] M. Otte, Size-selective molecular flasks, ACS Catal. 6 (2016) 6491–6510. 3619–3620.
[13] P.W.N.M. van Leeuwen, Supramolecular Catalysis, Wiley-VCH Verlag GmbH & Co. [45] C. Klöck, R.N. Dsouza, W.M. Nau, Cucurbituril-mediated supramolecular acid cat-
KGaA, 2008. alysis, Org. Lett. 11 (2009) 2595–2598.
[14] U.H. Brinker, J.-L. Mieusset, Molecular Encapsulation: Organic Reactions in [46] N. Basilio, L. García-Río, J.A. Moreira, M. Pessêgo, Supramolecular catalysis by
Constrained Systems, John Wiley & Sons, 2010. cucurbit[7]uril and cyclodextrins: similarity and differences, J. Org. Chem. 75
[15] S. Sadjadi, Organic Nanoreactors: From Molecular to Supramolecular Organic (2010) 848–855.
Compounds, Academic Press, 2016. [47] Y.-H. Wang, H. Cong, F.-F. Zhao, S.-F. Xue, Z. Tao, Q.-J. Zhu, G. Wei, Selective
[16] J. Kang, J. Rebek Jr., Acceleration of a Diels-Alder reaction by a self-assembled catalysis for the oxidation of alcohols to aldehydes in the presence of cucurbit[8]
molecular capsule, Nature 385 (1997) 50–52. uril, Catal. Commun. 12 (2011) 1127–1130.
[17] T. Iwasawa, R.J. Hooley, J. Rebek Jr., Stabilization of labile carbonyl addition in- [48] H. Cong, F.-F. Zhao, J.X. Zhang, X. Zeng, Z. Tao, S.-F. Xue, Q.-J. Zhu, Rapid
termediates by a synthetic receptor, Science 317 (2007) 493–496. transformation of benzylic alcohols to aldehyde in the presence of cucurbit[8]uril,
[18] M. Yoshizawa, M. Tamura, M. Fujita, Diels-Alder in aqueous molecular hosts: Catal. Commun. 11 (2009) 167–170.
unusual regioselectivity and efficient catalysis, Science 312 (2006) 251–254. [49] X. Lu, E. Masson, Silver-promoted desilylation catalyzed by ortho- and allosteric
[19] D.M. Kaphan, M.D. Levin, R.G. Bergman, K.N.F. Raymond, D. Toste, A supramo- cucurbiturils, Org. Lett. 12 (2010) 2310–2313.
lecular microenvironment strategy for transition metal catalysis, Science 350 [50] L. Scorsin, J.A. Roehrs, R.R. Campedelli, G.F. Caramori, A.O. Ortolan,
(2015) 1235–1238. R.L.T. Parreira, H.D. Fiedler, A. Acuña, L. García-Río, F. Nome, Cucurbituril-
[20] L. Xu, S. Hua, S. Li, Insight into the reaction between a primary amine and a ca- mediated catalytic hydrolysis: a kinetic and computational study with neutral and
vitand with an introverted aldehyde group: an enzyme-like mechanism, Chem. cationic dioxolanes in CB7, ACS Catal. 8 (2018) 12067–12079.
Commun. 49 (2013) 1542–1544. [51] S.M. de Lima, J.A. Gómez, V.P. Barros, G. de S. Vertuan, M. das D. Assis, C.F. de O.
[21] L. Xu, W. Hua, S. Hua, J. Li, S. Li, Mechanistic insight on the Diels−Alder reaction Graeff, G.J.-F. Demets, A new oxovanadium(IV)-cucurbit[6]uril complex: properties
catalyzed by a self-assembled molecular capsule, J. Org. Chem. 78 (2013) and potential for confined heterogeneous catalytic oxidation reactions, Polyhedron
3577–3582. 29 (2010) 3008–3013.
[22] L. Xu, G. Fang, S. Li, Supramolecular catalysis in the methylation of meta-phenylene [52] S.M. Bruno, A.C. Gomes, T.S.M. Oliveira, M.M. Antunes, A.D. Lopes, A.A. Valente,
ethynylene foldamer containing N, N-dimethylaminopyridine, RSC Adv. 7 (2017) I.S. Gonçalves, M. Pillinger, Catalytic alcoholysis of epoxides using metal-free cu-
14046–14052. curbituril-based solids, Org. Biomol. Chem. 14 (2016) 3873–3877.
[23] L. Xu, G. Fang, J. Tao, Z. Ye, S. Xu, Z. Li, Molecular mechanism and solvation effect [53] S. Das, T. Asefa, Epoxide ring-opening reactions with mesoporous silica-supported
of supramolecular catalysis in a synthetic cavitand receptor with an inwardly di- Fe(III) catalysts, ACS Catal. 1 (2011) 502–510.
rected carboxylic acid for ring-opening cyclization of epoxy alcohols, ACS Catal. 8 [54] K.W. Armbrust, M.G. Beaver, T.F. Jamison, Rhodium-catalyzed endo-selective ep-
(2018) 11910–11925. oxide-opening cascades: formal synthesis of (-)-brevisin, J. Am. Chem. Soc. 137
[24] T.-C. Lee, E. Kalenius, A.I. Lazar, K.I. Assaf, N. Kuhnert, C.H. Grün, J. Jänis, (2015) 6941–6946.
O.A. Scherman, W.M. Nau, Chemistry inside molecular containers in the gas phase, [55] E.N. Jacobsen, Asymmetric catalysis of epoxide ring-opening reactions, Acc. Chem.
Nat. Chem. 5 (2013) 376–382. Res. 33 (2000) 421–431.
[25] P. Carlqvista, F. Maseras, A theoretical analysis of a classic example of supramo- [56] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman,
lecular catalysis, Chem. Commun. 43 (2007) 748–750. G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato,
[26] C. Goehry, M. Besora, F. Maseras, Computational study on the mechanism of the X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M. Hada,
acceleration of 1,3-dipolar cycloaddition inside cucurbit[6]uril, ACS Catal. 5 (2015) M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda,
2445–2451. O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery Jr., J.E. Peralta, F. Ogliaro,
[27] S. He, F. Biedermann, N. Vankova, L. Zhechkov, T. Heine, R.E. Hoffman, M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, T. Keith,
A.D. Simone, T.T. Duignan, W.M. Nau, Cavitation energies can outperform dis- R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar,
persion interactions, Nat. Chem. 10 (2018) 1252–1257. J. Tomasi, M. Cossi, N. Rega, J.M. Millam, M. Klene, J.E. Knox, J.B. Cross,
[28] J. Lagona, P. Mukhopadhyay, S. Chakrabarti, L. Isaacs, The cucurbit[n]uril family, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev,
Angew. Chem. Int. Ed. 44 (2005) 4844–4870. A.J. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma,
[29] J.W. Lee, S. Samal, N. Selvapalam, H.-J. Kim, K. Kim, Cucurbituril homologues and V.G. Zakrzewski, G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich,
derivatives: new opportunities in supramolecular chemistry, Acc. Chem. Res. 36 A.D. Daniels, O. Farkas, J.B. Foresman, J.V. Ortiz, J. Cioslowski, D.J. Fox, Gaussian
(2003) 621–630. 09, Revision D.01, Gaussian, Inc., Wallingford CT, 2013.
[30] K. Kim, N. Selvapalam, Y.H. Ko, K.M. Park, D. Kim, J. Kim, Functionalized cu- [57] P.C. Hariharan, J.A. Pople, The influence of polarization functions on molecular
curbiturils and their applications, Chem. Soc. Rev. 36 (2007) 267–279. orbital hydrogenation energies, Theoret. Chimica Acta 28 (1973) 213–222.
[31] B.C. Pemberton, R. Raghunathan, S. Volla, J. Sivaguru, From containers to catalysts: [58] Y. Zhao, D.G. Truhlar, The M06 suite of density functionals for main group

7
L. Xu et al. Molecular Catalysis 467 (2019) 1–8

thermochemistry, thermochemical kinetics, noncovalent interactions, excited [65] K.P.C. Vollhardt, N.E. Schore, Organic Chemistry: Structure and Function, 6th
states, and transition elements: two new functionals and systematic testing of four edition, W. H. Freeman & Company, 2010.
M06-class functionals and 12 other functionals, Theor. Chem. Acc. 120 (2007) [66] L.G. Wade Jr., Organic Chemistry, 7th edition, Prentice Hall, 2009.
215–241. [67] R.L. Burwell Jr, R.G. Pearson, The principle of microscopic reversibility, J. Phys.
[59] Y. Zhao, D.G. Truhlar, Density functionals with broad applicability in chemistry, Chem. 70 (1966) 300–302.
Acc. Chem. Res. 41 (2008) 157–167. [68] P.L. Houston, Chemical Kinetics and Reaction Dynamics, Dover Publications Inc.,
[60] S. Grimme, J. Antony, S. Ehrlich, H. Krieg, A consistent and accurate ab initio Mineola, New York, 2001.
parametrization of density functional dispersion correction (DFT-D) for the 94 [69] I. Hwang, W.S. Jeon, H.-J. Kim, D. Kim, H. Kim, N. Selvapalam, N. Fujita,
elements H-Pu, J. Chem. Phys. 132 (2010) 154104. S. Shinkai, K. Kim, Cucurbit[7]uril: a simple macrocyclic, pH-triggered hydro-
[61] S. Grimme, Density functional theory with London dispersion corrections, WIREs gelator exhibiting guest-induced stimuli-responsive behavior, Angew. Chem. Int.
Comput. Mol. Sci. 1 (2011) 211–228. Ed. 46 (2007) 210–213.
[62] R.F. Ribeiro, A.V. Marenich, C.J. Cramer, D.G. Truhlar, Use of solution-phase vi- [70] S. Lim, H. Kim, N. Selvapalam, K.-J. Kim, S.J. Cho, G. Seo, K. Kim, Cucurbit[6]uril:
brational frequencies in continuum models for the free energy of solvation, J. Phys. organic molecular porous material with permanent porosity, exceptional stability,
Chem. B 115 (2011) 14556–14562. and acetylene sorption properties, Angew. Chem. Int. Ed. 47 (2008) 3352–3355.
[63] A.V. Marenich, C.J. Cramer, D.G. Truhlar, Universal solvation model based on so- [71] M. Yoon, K. Suh, H. Kim, Y. Kim, N. Selvapalam, K. Kim, High and highly aniso-
lute electron density and a continuum model of the solvent defined by the bulk tropic proton conductivity in organic molecular porous materials, Angew. Chem.
dielectric constant and atomic surface tensions, J. Phys. Chem. B 113 (2009) Int. Ed. 50 (2011) 7870–7873.
6378–6396. [72] D. Bardelang, K.A. Udachin, D.M. Leek, J.C. Margeson, G. Chan, C.I. Ratcliffe,
[64] C. Gonzalez, H.B. Schlegel, An improved algorithm for reaction path following, J. J.A. Ripmeester, Cucurbit[n]urils (n = 5–8): A comprehensive solid state study,
Chem. Phys. 90 (1989) 2154–2159. Cryst. Growth Des. 11 (2011) 5598–5614.

You might also like