Cooperative Interplay of Brønsted Acid and Lewis Acid Sites in MIL101 (CR) For Cross-Dehydrogenative Coupling of C H Bonds

You might also like

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

www.acsami.

org Research Article

Cooperative Interplay of Brønsted Acid and Lewis Acid Sites in MIL-


101(Cr) for Cross-Dehydrogenative Coupling of C−H Bonds
Jingwen Chen, Yuanyuan Zhang, Xiaoling Chen, Siyun Dai, Zongbi Bao, Qiwei Yang, Qilong Ren,
and Zhiguo Zhang*
Cite This: ACS Appl. Mater. Interfaces 2021, 13, 10845−10854 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.
Downloaded via NATL CHEMICAL LABORATORY on June 14, 2021 at 07:17:11 (UTC).

ABSTRACT: Cross-dehydrogenative coupling (CDC) is an effective


tool for carbon−carbon bond formation in chemical synthesis. Herein,
we report a metal−organic framework (MOF) possessing dual Lewis
acidic Cr sites and sulfonic acid sites (MIL-101(Cr)−SO3H) as an
efficient catalytic material for direct cross-coupling of xanthene and
different nucleophiles using O2 as the oxidant. The highly porous
structure of MIL-101(Cr)−SO3H enables the free access of reactants to
the catalytic active sites inside MOF pores. Kinetic studies indicated that
the Cr sites of MOF accelerate the rate-limiting autoxidation reaction of
xanthene, which synergistically work with the sulfonic acid group on
MOF ligands in promoting the CDC reactions. Besides, the catalytic
system shows excellent functional group compatibility, and a variety of
valuable xanthene derivatives were synthesized with satisfactory yields. Furthermore, MIL-101(Cr)−SO3H can be reused and its
catalytic activity and crystal structure remain after six consecutive runs.
KEYWORDS: cross-dehydrogenative coupling, metal−organic framework, sulfonic acid, xanthene, oxygen, C−H activation

■ INTRODUCTION
The cross-dehydrogenative coupling (CDC) of two different
Xanthene structure widely exists in bioactive compounds,
dyes, fluorescent materials, and so forth.27−29 Various method-
ologies have been reported for the preparation of xanthene
carbon−hydrogen bonds is a powerful strategy for the
derivatives, particularly the synthesis of 9-substituted xanthene
synthesis of organic molecules, upon which new carbon− derivatives via oxidative CDC reaction between benzylic
carbon bonds are formed in an atom-economic pathway C(sp3 )−H of xanthene and various active methylene
without the tedious prefunctionalization of substrates.1−3 The compounds.30−34 Transition-metal salts and peroxides are
C−H bonds are universal in organic chemicals;4 however, the often used as close partners to activate carbon−hydrogen
inertness of C−H makes it difficult for direct activation.5 bonds for CDC reactions. Recently, Klussmann et al.
Therefore, it is desirable to exploit efficient catalytic systems demonstrated that methane sulfonic acid was effective for the
for CDC reactions between the unactivated C−H bonds. Li CDC reactions of xanthene and enolizable compounds with
and co-workers6,7 have pioneered in this field and developed a oxygen as the oxidant.35−37 Elevated pressure of oxygen was
number of catalytic methods for the construction of carbon− required in the case of nucleophiles with poor reactivity.38
carbon bonds by direct functionalization of carbon−hydrogen Additionally, notorious corrosion and potential environmental
bonds. In recent years, various transition metal catalysis8,9 and pollution restrict their practical applications. Hence, the
metal-free catalytic strategies10−12 have been developed for the development of effective solid acid catalysts for sustainable
CDC reactions. Generally, an appropriate sacrificial oxidant is synthesis of xanthene derivatives is of great significance. Taking
required to accept the hydrogen.13 On behalf of green and this in mind, Maggi and co-workers31 reported that the
sustainable chemistry, the abundant and environmentally commercially available Amberlyst-15 catalyzed the CDC
benign molecular oxygen has attracted great attention as an reaction of xanthene and thioxanthene with different
oxidant in the oxidative CDC reactions.14−19 Tremendous
progress has been achieved in this field recently; however, the Received: November 15, 2020
process still faces drawbacks of long reaction time, poor Accepted: February 19, 2021
functional group compatibility, harsh reaction conditions, and Published: March 2, 2021
low product selectivity.20−26 Therefore, the development of
efficient and environmentally friendly catalytic systems for the
aerobic CDC reaction is still a significant issue.

© 2021 American Chemical Society https://dx.doi.org/10.1021/acsami.0c20369


10845 ACS Appl. Mater. Interfaces 2021, 13, 10845−10854
ACS Applied Materials & Interfaces www.acsami.org Research Article

Scheme 1. MIL-101(Cr)−SO3H-Catalyzed Aerobic CDC Reactionsa

a
H, O, C, S, and Cr are in blue, red, light gray, yellow, and green, respectively.

Figure 1. (A) PXRD patterns and (B) N2 sorption isotherms of MIL-101(Cr), MIL-101(Cr)−SO3H, and Amberlyst-15 at 77 K. SEM images of
MIL-101(Cr) (C) and MIL-101(Cr)−SO3H (D).

methylene compounds, which are restricted to limited molecular oxygen as the oxidizing agent. Meanwhile, the
substrate scope. Hence, regardless of these advances, much autoxidation of benzylic C−H of xanthene is the rate-limiting
effort in search of more efficient and practical methods is still step in sulfonic acid-catalyzed cross-coupling reaction of
called for the synthesis of xanthene derivatives. xanthene. These pioneer works inspired us to design a solid
Metal−organic frameworks (MOFs), prepared from the self- MOF catalyst with Cr centers and sulfonic acid sites,62,63
assembly of metal clusters/ions and organic ligands, have which possess dual active sites for accelerating the autoxidation
attracted remarkable research interest in catalysis, gas of xanthene and subsequent C−C coupling reactions to obtain
separation, sensing, and so forth.39−43 MOFs possess many 9-substituted xanthene derivatives. Following the above
advantages such as large surface area, well-defined structure, catalyst design criteria, in this study, we synthesized a
and available functionalization,44−46 and they have been an bifunctional, highly porous, and robust MOF, that is, MIL-
intriguing class of heterogeneous catalytic materials for a 101(Cr)−SO3H, for the CDC reaction of xanthene and
diverse range of organic transformations.47−51 Among various nucleophiles with O2 as the oxidizing agent (Scheme 1). High
MOFs reported, MIL-101 (Matérial Institut Lavoisier) with activity was observed for the MIL-101(Cr)−SO3H-catalyzed
two kinds of mesoporous cages (3.4 and 2.9 nm, respectively) CDC reaction between xanthene and cyclopentanone, which is
and high thermal and chemical stability has been a striking superior to the commercial solid acids. Kinetic studies
material for catalysis.52−54 The water molecules coordinated indicated that the Cr sites can accelerate the xanthene
on the chromium clusters can be eliminated upon heating in autoxidation efficiently and improve the reaction rate constant
vacuum, which generates coordinatively unsaturated Cr sites by 1.92 times. The cooperative action between Cr centers and
that can work as Lewis acid sites for catalysis.55−60 Very −SO3H groups of MIL-101(Cr)−SO3H is crucial to achieve
recently, Garcia et al. reported that MIL-101(Cr) with high catalytic activities. The catalytic system is compatible with
abundant Cr sites catalyzes the benzylic autoxidation of diverse ketones, and other substrates with easily oxidized
indane.61 A mechanism study revealed that MIL-101(Cr) acts groups such as aldehydes and inert malonates were well
as a radical initiator in the autoxidation of benzylic C−H with compatible with the reaction system as well. Moreover, with
10846 https://dx.doi.org/10.1021/acsami.0c20369
ACS Appl. Mater. Interfaces 2021, 13, 10845−10854
ACS Applied Materials & Interfaces www.acsami.org Research Article

Table 1. Catalytic Activities of Different Catalysts for the CDC Reactiona

entry catalyst H+ loading (mmol·g−1) BET surface area (m2·g−1) yieldb (%) selectivityc (%)
d
1 0 (5.2)
2 MIL-101(Cr)−SO3H 1.42 2362 63 98
3 Amberlyst-15 4.86 44 39 97
4 MIL-101(Cr)e 3517 0.4 (87)f 1
5 TsOH 5.81 82 91
6 MIL-101(Cr) + TsOHg 5.81 3517 83 94
a
Reaction conditions: xanthene (0.2 mmol), cyclopentanone (1.0 mmol), catalyst (5 mol % H+), O2 (1 bar), room temperature (r.t.), 28 h. bYields
were calculated by 1H NMR with dimethyl terephthalate as the internal standard. cSelectivity calculated based on xanthene. dYield of xanthone (4)
is given in parenthesis. e4.9 mg of MIL-101(Cr). fYield of oxidative byproducts. g4.9 mg of MIL-101(Cr) and 5 mol % of TsOH. TsOH = p-
toluenesulfonic acid.

the addition of a chiral imidazolidinone as the cocatalyst, (Figure 1D) are octahedral, and the particle size of MIL-
higher enantioselectivity was obtained with MIL-101(Cr)− 101(Cr)−SO3H is 200−700 nm with some extent of
SO3H as the catalyst than that with aromatic sulfonic acid, aggregation. Thermogravimetric analysis (TGA) of MIL-
suggesting the potential use of MOFs with chiral organo- 101(Cr)−SO3H was performed in air to evaluate the stability
catalysts for enantiomeric catalysis. of the material. A weight loss from 25 to 100 °C on the TGA


profile (Figure S1) attributes to the elimination of water
EXPERIMENTAL SECTION molecules, and the next one is caused by the removal of −OH
Synthesis of the Catalyst. MIL-101(Cr)−SO3H was prepared or F; the framework decomposes from 300 to 400 °C. The
and purified through a method described in the literature.64,65 In brief, TGA result revealed that MIL-101(Cr)−SO3H shows high
monosodium 2-sulfoterephthalate (5.4 g), Cr(NO3)3·9H2O (4.0 g), thermal stability (up to 250 °C).
DI water (60 mL), and hydrofluoric acid (37 wt %, 520 μL) were Catalytic CDC Reactions of Xanthene and Nucleo-
mixed in a Teflon-lined stainless steel autoclave. The mixture was philes. The CDC reaction of xanthene and cyclopentanone
heated to 190 °C. After 24 h, the mixture was cooled to room was chosen to evaluate the catalytic activity of solid acids. The
temperature, and the resulting solid was washed with H2O and H+ loading of the catalyst was measured by a titration method
methanol (×5) and collected by centrifugation. The solid was dried reported in the literature.66 As shown in Table 1, almost no
under vacuum at 120 °C before measuring the N2 sorption isotherms.
Catalytic CDC Reaction of Xanthene with Nucleophiles. In
reaction occurs even after 28 h without the catalyst (entry 1).
general, xanthene (0.2 mmol), nucleophiles (5.0 equiv), solvent, and The CDC reaction product (3a) was obtained in 63% yield
the catalyst were introduced into a reaction tube. The reactor was after introducing MIL-101(Cr)−SO3H into the reaction
flushed with oxygen to eliminate the air. Then, the mixture was stirred system (entry 2). Under the same reaction conditions, the
at room temperature or in a preheated oil bath. A balloon filled with catalytic activity of Amberlyst-15 is much inferior, and only
oxygen was used to provide the desired gas atmosphere. After the 39% of 3a was formed (entry 3). The superior activity of MIL-
reaction, the product was separated by column chromatography with 101(Cr)−SO3H to Amberlyst-15 may originate from the high
hexane−dichloromethane as the eluent. All products’ structures were porosity of MOF that enables the easy accessibility of reactants
verified by NMR and mass spectra.


to the acid sites and the Cr sites. Hence, to investigate the
function of Cr sites in the CDC reaction, the catalytic activity
RESULTS AND DISCUSSION of MIL-101(Cr), an isostructural MOF to MIL-101(Cr)−
Synthesis and Characterization. MIL-101(Cr)−SO3H SO3H, was tested for the CDC reaction (entry 4). In the
was synthesized through a hydrothermal reaction of mono- presence of an equal amount of Cr sites (Table S1), only a
sodium 2-sulfoterephthalate and Cr(NO3)3·9H2O by following trace amount of 3a was detected, while a large amount of
a reported method with some modifications.64,65 The powder benzylic C−H oxidation product (4 + 6 + 7) was formed.
X-ray diffraction (PXRD) pattern of MIL-101(Cr)−SO3H Besides, p-toluenesulfonic acid (TsOH), an analogue of
matches well with that of simulated one (Figure 1A), organic linkers of MIL-101(Cr)−SO3H, was tested in
indicating that the insertion of sulfonic groups into the catalyzing the CDC reaction, 3a was obtained in 82% yield
skeleton of MIL-101(Cr) has no influence on the MOF (entry 5). To investigate the influence of Cr on the CDC
structure. N2 sorption isotherms of the solid catalysts measured reaction, a mixture of MIL-101(Cr) and TsOH was used as the
at 77 K are presented in Figure 1B. The Brunauer−Emmett− catalyst, and the yield of 3a was 83% (entry 6). Hence, the Cr
Teller (BET) surface area of MIL-101(Cr) is 3517 m2·g−1, sites of MIL-101(Cr)−SO3H do not improve the yield of 3a
which decreases to 2362 m2·g−1 after the incorporation of directly, and further studies are needed to gain deep insights
sulfonic acid groups inside the pore (Table S1). The pore into the catalytic mechanism of MIL-101(Cr)−SO3H.
volume of MIL-101(Cr) and MIL-101(Cr)−SO3H were 1.75 We further examined other reaction parameters on the
and 1.33 cm3·g−1, respectively. Amberlyst-15, a commercially reaction outcomes. As shown in Table 2, the addition of
available solid acid, exhibits a low BET surface area (46 m2· solvent causes detrimental effect on the reaction (entries 2−5).
g−1). Scanning electronic microscopy (SEM) images show that A yield of 63% was obtained at room temperature without
both MIL-101(Cr) (Figure 1C) and MIL-101(Cr)−SO3H solvent (entry 1). The yield was improved to 89% upon
10847 https://dx.doi.org/10.1021/acsami.0c20369
ACS Appl. Mater. Interfaces 2021, 13, 10845−10854
ACS Applied Materials & Interfaces www.acsami.org Research Article

Table 2. Effect of Other Reaction Parameters on the Under the optimized reaction conditions, we then
Aerobic Oxidative CDC Reactiona investigated the MIL-101(Cr)−SO3H-catalyzed CDC reac-
tions between xanthene (1) and various nucleophiles (2)
entry solvent yieldb (%)
(Figure 2). For cyclic ketones (2a−2c, 2k), the products were
1 63 formed in good yields. However, the yield decreased slightly
2 CH2Cl2 13 with large size substrates (2c and 2k). The acyclic ketones
3 toluene 17 were also well compatible with the mild reaction conditions
4 hexane 10 (3d−3h). For 1′-acetonaphthone (2i) and 9-acetylanthracene
5 EtOAc 36 (2j), a small amount of ethyl acetate was introduced to
6c 18 dissolve the solid substrates. The product (3i and 3j) yield
7d 89
a
decreased obviously, which was probably caused by the
Reaction conditions: xanthene (0.2 mmol), cyclopentanone (1 detrimental effect of the solvent (Table 2). With β-keto esters
mmol), MIL-101(Cr)−SO3H (7 mg, 5 mol % of H+), solvent (1 mL), as the nucleophiles (3l−3n), a moderate yield was obtained.
O2 (1 bar), r.t., 28 h. bNMR yield. c1 bar of air. d40 °C.
Remarkably, malonate, an inert substrate even under high-
pressure O2,36 was engaged successfully in the CDC reactions
with 64% yield (3o) under the catalysis of MIL-101(Cr)−
elevating the reaction temperature to 40 °C (entry 7). Using SO3H in oxygen flow. Aldehydes could also be employed as
air as the oxidant (entry 6), the yield of 3a decreased the nucleophiles when the catalyst amount was changed to 21
dramatically to 18%. A survey of catalyst loading shows that 7 mg in CH3NO2 under 1 bar of O2 at room temperature for 3 d
mg of MIL-101(Cr)−SO3H is the optimum amount (Figure (Figure S3). It is worthy to note that the overoxidation of the
S2). aldehyde group to the corresponding acids was inhibited in the

Figure 2. Scope of the MIL-101(Cr)−SO3H-catalyzed CDC reaction of xanthene and nucleophiles. a Reaction conditions: xanthene (0.2 mmol), 2
(1.0 mmol), MIL-101(Cr)−SO3H (7 mg), O2 (1 bar), 40 °C, 24 h. Yield of isolated product. b 0.2 mL of EtOAc was used as the solvent. c The
reaction was conducted with 1 mmol scale, 60 °C, continuous O2 flow (3.0 mL·min−1), 2 d. d CH3NO2 (1.0 mL), MIL-101(Cr)−SO3H (21 mg),
r.t., 3 d.

10848 https://dx.doi.org/10.1021/acsami.0c20369
ACS Appl. Mater. Interfaces 2021, 13, 10845−10854
ACS Applied Materials & Interfaces www.acsami.org Research Article

presence of MIL-101(Cr)−SO3H (3p−3t). In contrast, when (Figure S5), albeit the particle sizes decreased, as shown by the
Amberlyst-15 was used as the catalyst, only a trace amount of SEM image (Figure S6).
the cross-coupling product was formed from xanthene and Catalytic Mechanism. The high activity of MIL-
aldehyde. Therefore, MIL-101(Cr)−SO3H has been proved as 101(Cr)−SO3H in the CDC reaction for the synthesis of
an effective and versatile solid catalyst for CDC reactions. xanthene derivatives prompted us to put some deep insights
After having investigated the scope of nucleophiles, we into the catalytic mechanism. It was presumed that the CDC
turned to study the reactivity of other “electrophiles” besides reaction proceeded through the autoxidation of xanthene via a
xanthene. Expanding the CDC reaction substrate scope to hydroperoxide intermediate.38 Here, we carried out a series of
other benzylic substrates such as thioxanthene (2u) and N- kinetic studies to evaluate the role of active sites in MIL-
methylacridane (2v) is another challenge, and high oxygen 101(Cr)−SO3H and the catalytic reaction mechanism.
pressure (5−10 bar) and long reaction time are required, as Initially, xanthene and cyclopentanone were mixed and stirred
reported in previous work.35,36 To our delight, these benzylic continuously at 40 °C under O2 without the catalyst. As shown
substrates can react with cyclopentanone under ambient in Figure 4A, no CDC product (3a) was observed, while
oxygen pressure, and the CDC reaction products were formed xanthene hydroperoxide 6 was formed as the major product.
in 41% and 15% yield, respectively (Figure 3). Other oxidation byproducts (4−7), meanwhile, were observed
during the reaction process (Figure 4). After 5 h, MIL-
101(Cr)−SO3H was introduced into the reaction mixture. The
cross-coupling product 3a was formed immediately accom-
panied by the rapid disappearance of xanthene hydroperoxide.
The concentrations of xanthone (4) and xanthydrol (5) were
low during the whole reaction process. This result illustrated
that hydroperoxide species are the reactive intermediates for
the CDC reaction.
The high BET surface area and large pore sizes of MIL-
101(Cr)−SO 3H enable the catalytic active sites fully
accessible, which contribute to its high catalytic activity.
Besides, the high density of Cr sites in MIL-101(Cr)−SO3H is
quite different from that of Amberlyst-15, which may affect the
Figure 3. Scope of the MIL-101(Cr)−SO3H-catalyzed CDC reaction
catalytic activity. To investigate the function of Cr sites, the
of benzylic hydrocarbons and cyclopentanone. Reaction conditions: reaction between xanthene and cyclopentanone was conducted
MIL-101(Cr)−SO3H (5 mol % of H+), O2 (1 bar), 40 °C, 56 h. Yield with MIL-101(Cr) as the catalyst to exclude the influence of
of the product isolated by column chromatography. sulfonic acid groups. A control experiment without the catalyst
was also carried out for comparison. As shown in Figure 5A,
xanthene hydroperoxide 6 can be generated continuously
without the catalyst. After 5 h, 6 gradually decomposed and
To check if the CDC reaction proceeds heterogeneously, we converted to xanthone and xanthydrol. After the addition of
conducted a filtration experiment. After 7 h, the catalyst was MIL-101(Cr), the conversion of xanthene was accelerated
removed by filtration. No increase of the coupling product obviously (Figure 5B). The time-resolved data of xanthene was
yield was observed (Table S2), which suggested that MIL- curve-fitted by pseudo-first-/second-order reaction rate equa-
101(Cr)−SO3H catalyzed the reaction heterogeneously. The tions. As shown in Figure 6B, the autoxidation of xanthene
reusability of the catalyst was tested by the CDC reaction behaves as a first-order process. The rate constant of xanthene
between xanthene and cyclopentanone. The yield and under the catalysis of MIL-101(Cr) is 0.3107 h−1 (Figure 6B),
selectivity of 3a remained almost unchanged after six which is 1.92 times higher than that without the catalyst (0.162
consecutive runs (Figure S4). The crystal structure of MIL- h−1, Figure 6A). Note that hydroperoxide 6 decomposed more
101(Cr)−SO3H was kept as indicated by PXRD patterns rapidly and transformed to the oxidation byproduct (xanthone,

Figure 4. (A) Kinetic profiles for the reaction between xanthene and cyclopentanone. Reaction conditions: xanthene (1.0 mmol), cyclopentanone
(5.0 mmol), O2 (1 bar), 40 °C. Adding MIL-101(Cr)−SO3H (7 mg) at a reaction time of 5 h. (B) CDC reaction of xanthene (1) and
cyclopentanone (2a) for the preparation of 3a and structures of related intermediates and byproducts (4−7).

10849 https://dx.doi.org/10.1021/acsami.0c20369
ACS Appl. Mater. Interfaces 2021, 13, 10845−10854
ACS Applied Materials & Interfaces www.acsami.org Research Article

Figure 5. Composition of the reaction mixture of xanthene and cyclopentanone without catalyst (A) and with MIL-101(Cr) (B). Reaction
conditions: xanthene (1.0 mmol), cyclopentanone (5.0 mmol), O2 (1 bar), 40 °C, 7 mg of MIL-101(Cr) was introduced in (B).

Figure 6. Kinetic profiles of xanthene autoxidation reaction fitted by first- and second-order reaction rate equations. Reaction conditions: xanthene
(1.0 mmol), cyclopentanone (5.0 mmol), O2 (1 bar), 40 °C (A), and 7 mg of MIL-101(Cr) was introduced in (B).

Scheme 2. Plausible Reaction Mechanism of the MIL-101(Cr)−SO3H-Catalyzed CDC Reaction between Xanthene and
Cyclopentanone

4) in the presence of MIL-101(Cr). Therefore, MIL-101(Cr) nucleophiles to form 3a. Hence, the coexistence and relay
accelerated the rate-limiting aerobic autoxidation of xanthene catalysis of Cr sites and sulfonic acid groups in MIL-101(Cr)−
and the decomposition and conversion of intermediate 6. SO3H are pivotal for improving the CDC reaction rate and
However, when introducing sulfonic acid groups into MIL- product selectivity. The existence of peroxide intermediate 6
101(Cr), the generated hydroperoxide 6 turned to couple with and 7 was further verified by mass spectrometry. In the
10850 https://dx.doi.org/10.1021/acsami.0c20369
ACS Appl. Mater. Interfaces 2021, 13, 10845−10854
ACS Applied Materials & Interfaces www.acsami.org Research Article

Table 3. Asymmetric CDC Reaction of Xanthene and Pentanal under Different Conditionsa

a
Reaction conditions: xanthene (0.2 mmol), pentanal (1 mmol), catalyst (15 mol % in terms of H+ amount), CH3NO2 (1 mL), O2 (1 bar), 5 °C, 5
d. bYield determined by 1H NMR. cee value was analyzed by chiral HPLC analysis.

reaction mixture, peaks at m/z = 237.0 and 417.1 (ESI) were 101(Cr)−SO3H, carbon cation 10 and H2O2 were formed.
observed by electrospray ionization mass spectrometry (ESI- Finally, the enol form 11, which results from the interaction of
MS) under the catalysis of MIL-101(Cr). However, when 2 with the Cr sites, attacks carbocation 10 via the SN1-type
using MIL-101(Cr)−SO3H as the catalyst, these peroxide reaction to give the cross-coupling product 3. The MIL-
intermediates were hardly observed due to the fast reaction of 101(Cr)−SO3H catalyst is regenerated and goes to the next
the peroxide intermediate with cyclopentanone. As a result, we catalytic cycle. The cooperation between Cr centers and
conclude that the synergistic effect of Cr centers and −SO3H −SO3H groups of MIL-101(Cr)−SO3H makes it an excellent
sites in MIL-101(Cr)−SO3H is the main reason for its superior catalyst for the CDC reaction of two carbon−hydrogen bonds.
catalytic performance over the commercial solid acid catalysts After the success in developing the MOF-catalyzed CDC
such as Amberlyst-15. reaction of C−H bonds, we speculate whether an enantiose-
The hydroperoxide intermediate 6 reacted with nucleophiles lective CDC reaction via C−H activation could be realized.
to generate cross-coupling products, and hydrogen peroxide Recently, Cozzi’s67 and Jiao’s32 groups have reported the
was expected to be formed as well. To verify the existence of asymmetric reaction between xanthene and aldehydes
H2O2, we used sodium iodide as an indicator. No obvious promoted by MacMillan-type organocatalysts through enamine
color change was observed for the pure sodium iodide solution catalysis and acid catalysis. We envisioned that this type of
under air even after 12 h. However, the colorless solution of catalytic mode could be realized by the combination of an
sodium iodide turned to yellow immediately upon adding the enamine catalyst with Brønsted acid-functionalized MOFs. To
filtrate of MIL-101(Cr)−SO3H-catalyzed CDC reaction validate our hypothesis, we introduced a chiral imidazolidinone
mixture (Figure S7), which illustrated that hydrogen peroxide to the reaction mixture of xanthene and pentanal catalyzed by
was generated in the cross-coupling step. The above kinetic MIL-101(Cr)−SO3H. Table 3 shows that the desired CDC
studies and color change experiment demonstrated that the reaction product 3q was obtained with 78% enantiomeric
CDC reaction proceeded via the peroxide species, but it is not excess (ee) with MIL-101(Cr)−SO3H as the catalyst (entry 2),
clear how the peroxide intermediates formed. To gain clear which is much superior to that of the organic sulfonic acid
understanding of the reaction process, we added 3 equivalents catalyst (entry 1, 24% ee). The improved ee might originate
of 2,6-di-tert-butyl-4-methyl phenol (BHT) as the radical trap from the confinement effect of MOF pores.68 In addition, the
in the cross-coupling reaction mixture (Table S4, entry 1), only ee value increased as more chiral imidazolidinone was added
2% of the CDC product was observed. However, if xanthene and reached an 87% ee with 25 mol % of the cocatalyst (entry
hydroperoxide was used as the starting material, the desired 4), underlining the potential use of MOFs with chiral
product was formed without being affected by BHT (Table S4, organocatalysts in enantiomeric synthesis of biologically active
chiral molecules.


entry 2). Hence, the formation of the xanthene peroxide
intermediate involves a radical pathway.
Based on previous studies38,61 and the above experiment, a CONCLUSIONS
plausible catalytic mechanism was put forward (Scheme 2). In summary, a Cr- and sulfonic acid-functionalized MOF-
The Cr clusters in MIL-101(Cr)−SO3H interact with oxygen catalyzed aerobic CDC reaction of two carbon−hydrogen
to form metal-superoxo species, which work as the radical bonds was developed. The large surface area and pore size
initiator to accelerate the autoxidation of xanthene (1). Then, combined with bifunctional catalytic active sites of MIL-
the Cr metalloperoxides abstract a hydrogen from the benzylic 101(Cr)−SO3H have demonstrated excellent catalytic per-
C−H bond of xanthene and generate radical 8, which is formance toward CDC reactions, and a diverse range of
immediately trapped by O2 and forms hydroperoxide 6 via xanthene derivatives were synthesized in moderate to high
radical 9. Then, with the assistance of sulfonic acid from MIL- yields. Various easily oxidized aldehydes are tolerant to these
10851 https://dx.doi.org/10.1021/acsami.0c20369
ACS Appl. Mater. Interfaces 2021, 13, 10845−10854
ACS Applied Materials & Interfaces www.acsami.org Research Article

reaction conditions. Furthermore, the inert malonate can react Qilong Ren − Key Laboratory of Biomass Chemical
smoothly with xanthene catalyzed by MIL-101(Cr)−SO3H Engineering of Ministry of Education, College of Chemical
when switching oxygen balloon to oxygen flow. A thorough and Biological Engineering, Zhejiang University, Hangzhou
kinetic study reveals that the Cr sites of MIL-101(Cr)−SO3H 310027, P. R. China; Institute of Zhejiang University
accelerate the rate-limiting aerobic autoxidation of xanthene. Quzhou, Quzhou 324000, P. R.China
Moreover, with the addition of a chiral organocatalyst as the Complete contact information is available at:
cocatalyst, chiral xanthene derivatives with up to 87% ee were https://pubs.acs.org/10.1021/acsami.0c20369
obtained. These discoveries highlight the potential of multi-
functional MOFs as facile and efficient catalysts for aerobic Notes
CDC reactions to synthesize valuable bioactive molecules.
The authors declare no competing financial interest.


*
ASSOCIATED CONTENT
sı Supporting Information
■ ACKNOWLEDGMENTS
This work was supported by the National Key R&D Program
The Supporting Information is available free of charge at of China (grant number 2016YFA0202900), the National
https://pubs.acs.org/doi/10.1021/acsami.0c20369. Natural Science Foundation of China (grant numbers
Pore structure and TGA profiles of catalysts; catalyst 21878266 and 22078288), and China Postdoctoral Science
loading effect; reaction kinetics; recycling of MIL- Foundation (2020M680077).
101(Cr)−SO3H; and characterization data of the CDC
products (PDF) ■ REFERENCES


(1) Huang, C.-Y.; Kang, H.; Li, J.; Li, C.-J. En Route to
Intermolecular Cross-Dehydrogenative Coupling Reactions. J. Org.
AUTHOR INFORMATION Chem. 2019, 84, 12705−12721.
Corresponding Author (2) Scheuermann, C. J. Beyond Traditional Cross Couplings: The
Zhiguo Zhang − Key Laboratory of Biomass Chemical Scope of the Cross Dehydrogenative Coupling Reaction. Chem. -
Asian J. 2010, 5, 436−451.
Engineering of Ministry of Education, College of Chemical
(3) Li, C.-J. Cross-Dehydrogenative Coupling (CDC): Exploring
and Biological Engineering, Zhejiang University, Hangzhou C−C Bond Formations beyond Functional Group Transformations.
310027, P. R. China; Institute of Zhejiang University Acc. Chem. Res. 2009, 42, 335−344.
Quzhou, Quzhou 324000, P. R.China; orcid.org/0000- (4) Girard, S. A.; Knauber, T.; Li, C.-J. The Cross-Dehydrogenative
0003-1681-4853; Email: zhiguo.zhang@zju.edu.cn Coupling of Csp3−H Bonds: A Versatile Strategy for C−C Bond
Formations. Angew. Chem., Int. Ed. 2014, 53, 74−100.
Authors (5) Bergman, R. G. C-H activation. Nature 2007, 446, 391−393.
Jingwen Chen − Key Laboratory of Biomass Chemical (6) Li, Z.; Li, C.-J. Highly Efficient Copper-Catalyzed Nitro-
Engineering of Ministry of Education, College of Chemical Mannich Type Reaction: Cross-Dehydrogenative-Coupling between
and Biological Engineering, Zhejiang University, Hangzhou sp3 C−H Bond and sp3 C−H Bond. J. Am. Chem. Soc. 2005, 127,
310027, P. R. China; Institute of Zhejiang University 3672−3673.
Quzhou, Quzhou 324000, P. R.China; orcid.org/0000- (7) Gong, H.; Zeng, H.; Zhou, F.; Li, C.-J. Rhodium(I)-Catalyzed
Regiospecific Dimerization of Aromatic Acids: Two Direct C−H
0001-6724-5414 Bond Activations in Water. Angew. Chem., Int. Ed. 2015, 54, 5718−
Yuanyuan Zhang − Key Laboratory of Biomass Chemical 5721.
Engineering of Ministry of Education, College of Chemical (8) Li, X.; Ouyang, W.; Nie, J.; Ji, S.; Chen, Q.; Huo, Y. Recent
and Biological Engineering, Zhejiang University, Hangzhou Development on Cp*Ir(III)-Catalyzed C-H Bond Functionalization.
310027, P. R. China; Institute of Zhejiang University ChemCatChem 2020, 12, 2358−2384.
Quzhou, Quzhou 324000, P. R.China (9) Shang, X.; Liu, Z.-Q. Transition Metal-Catalyzed Cvinyl−Cvinyl
Xiaoling Chen − Key Laboratory of Biomass Chemical Bond Formation via Double Cvinyl−H Bond Activation. Chem. Soc.
Engineering of Ministry of Education, College of Chemical Rev. 2013, 42, 3253−3260.
and Biological Engineering, Zhejiang University, Hangzhou (10) Parvatkar, P. T.; Manetsch, R.; Banik, B. K. Metal-Free Cross-
310027, P. R. China Dehydrogenative Coupling (CDC): Molecular Iodine as a Versatile
Catalyst/Reagent for CDC Reactions. Chem.Asian J. 2019, 14, 6−
Siyun Dai − Key Laboratory of Biomass Chemical Engineering 30.
of Ministry of Education, College of Chemical and Biological (11) Kouznetsov, V. V.; Ortiz-Villamizar, M. C.; Méndez-Vargas, L.
Engineering, Zhejiang University, Hangzhou 310027, P. R. Y.; Galvis, C. E. P. A Review on Metal-Free Oxidative alpha-
China Cyanation and Alkynylation of N-Substituted Tetrahydroisoquino-
Zongbi Bao − Key Laboratory of Biomass Chemical lines as a Rapid Route for the Synthesis of Isoquinoline Alkaloids.
Engineering of Ministry of Education, College of Chemical Curr. Org. Chem. 2020, 24, 809−816.
and Biological Engineering, Zhejiang University, Hangzhou (12) Batra, A.; Singh, K. N. Recent Developments in Transition
310027, P. R. China; Institute of Zhejiang University Metal-Free Cross-Dehydrogenative Coupling Reactions for C-C Bond
Quzhou, Quzhou 324000, P. R.China; orcid.org/0000- Formation. Eur. J. Org. Chem. 2020, 6676−6703.
0003-4327-3028 (13) Yeung, C. S.; Dong, V. M. Catalytic Dehydrogenative Cross-
Coupling: Forming Carbon−Carbon Bonds by Oxidizing Two
Qiwei Yang − Key Laboratory of Biomass Chemical Carbon−Hydrogen Bonds. Chem. Rev. 2011, 111, 1215−1292.
Engineering of Ministry of Education, College of Chemical (14) Xu, D.-Z.; Hu, R.-M.; Lai, Y.-H. Iron-Catalyzed Aerobic
and Biological Engineering, Zhejiang University, Hangzhou Oxidative Cross-Dehydrogenative C(sp3)H/X-H (X = C, N, S)
310027, P. R. China; Institute of Zhejiang University Coupling Reactions. Synlett 2020, 31, 1753−1759.
Quzhou, Quzhou 324000, P. R.China; orcid.org/0000- (15) Sarma, D.; Majumdar, B.; Sarma, T. K. Visible-light induced
0002-6469-5126 enhancement in the multi-catalytic activity of sulfated carbon dots for

10852 https://dx.doi.org/10.1021/acsami.0c20369
ACS Appl. Mater. Interfaces 2021, 13, 10845−10854
ACS Applied Materials & Interfaces www.acsami.org Research Article

aerobic carbon-carbon bond formation. Green Chem. 2019, 21, 6717− (34) Li, Y.; Li, Y.; Li, Y.; Chen, C.; Ying, F.; Dong, Y.; Liang, D.
6726. Metal-Free Cross-Dehydrogenative C-N Coupling of Azoles with
(16) Rana, P.; Gaur, R.; Gupta, R.; Arora, G.; Jayashree, A.; Sharma, Xanthenes and Related Activated Arylmethylenes. Synth. Commun.
R. K. Cross-Dehydrogenative C(sp3)-C(sp3) Coupling via C-H 2019, 49, 2053−2065.
Activation Using Magnetically Retrievable Ruthenium-Based Photo- (35) Pintér, Á .; Sud, A.; Sureshkumar, D.; Klussmann, M.
redox Nanocatalyst under Aerobic Conditions. Chem. Commun. 2019, Autoxidative Carbon-Carbon Bond Formation from Carbon-Hydro-
55, 7402−7405. gen Bonds. Angew. Chem., Int. Ed. 2010, 49, 5004−5007.
(17) Wu, H.; Su, C.; Tandiana, R.; Liu, C.; Qiu, C.; Bao, Y.; Wu, J. (36) Pintér, Á .; Klussmann, M. Sulfonic Acid-Catalyzed Autoxidative
E.; Xu, Y.; Lu, J.; Fan, D.; Loh, K. P. Graphene-Oxide-Catalyzed Carbon-Carbon Coupling Reaction under Elevated Partial Pressure of
Direct CH-CH-Type Cross-Coupling: The Intrinsic Catalytic Oxygen. Adv. Synth. Catal. 2012, 354, 701−711.
Activities of Zigzag Edges. Angew. Chem., Int. Ed. 2018, 57, 10848− (37) Klussmann, M.; Schweitzer-Chaput, B. Interplay of Method
10853. Development and Mechanistic Studies−From Aerobic Oxidative
(18) Zhou, K.; Yu, Y.; Lin, Y.-M.; Li, Y.; Gong, L. Copper-Catalyzed Coupling to Radical Reactions via Alkenyl Peroxides. Synlett 2016, 27,
Aerobic Asymmetric Cross-Dehydrogenative Coupling of C(sp3)-H 190−202.
Bonds Driven by Visible Light. Green Chem. 2020, 22, 4597−4603. (38) Schweitzer-Chaput, B.; Sud, A.; Pintér, Á .; Dehn, S.; Schulze,
(19) Zhou, W.; Lu, W.; Wang, H.; Xia, Z.; Zhai, S.; Zhang, Z.; Ma, P.; Klussmann, M. Synergistic Effect of Ketone and Hydroperoxide in
Y.; He, M.; Chen, Q. CuMgAl Hydrotalcite as an Efficient Brønsted Acid Catalyzed Oxidative Coupling Reactions. Angew.
Bifunctional Catalyst for the Cross-Dehydrogenative C-C Coupling Chem., Int. Ed. 2013, 52, 13228−13232.
Reactions under Mild Conditions. Appl. Catal., A 2020, 604, 117771. (39) Ryu, U.; Jee, S.; Rao, P. C.; Shin, J.; Ko, C.; Yoon, M.; Park, K.
(20) Li, C.-J.; Liu, W.; Zhu, Y. Pd-Catalyzed Homo Cross- S.; Choi, K. M. Recent Advances in Process Engineering and
Dehydrogenative Coupling of 2-Arylpyridines by Using I2 as the Upcoming Applications of Metal-Organic Frameworks. Coord. Chem.
Sole Oxidant. Synthesis 2016, 48, 1616−1621. Rev. 2021, 426, 213544.
(21) Yazaki, R.; Ohshima, T. Recent Strategic Advances for the (40) Inamuddin; Boddula, R.; Ahamed, M. I.; Asiri, A. M.,
Activation of Benzylic C-H Bonds for the Formation of C-C Bonds. Applications of Metal−Organic Frameworks and Their Derived
Tetrahedron Lett. 2019, 60, 151225. Materials; John Wiley & Sons: Hoboken, 2020.
(22) Bagdi, A. K.; Rahman, M.; Bhattacherjee, D.; Zyryanov, G. V.; (41) Cai, X.; Xie, Z.; Li, D.; Kassymova, M.; Zang, S.-Q.; Jiang, H.-L.
Ghosh, S.; Chupakhin, O. N.; Hajra, A. Visible Light Promoted Cross- Nano-Sized Metal-Organic Frameworks: Synthesis and Applications.
Dehydrogenative Coupling: a Decade Update. Green Chem. 2020, 22, Coord. Chem. Rev. 2020, 417, 213366.
6632−6681. (42) Kreno, L. E.; Leong, K.; Farha, O. K.; Allendorf, M.; Van
(23) Bosque, I.; Chinchilla, R.; Gonzalez-Gomez, J. C.; Guijarro, D.; Duyne, R. P.; Hupp, J. T. Metal-Organic Framework Materials as
Alonso, F. Cross-Dehydrogenative Coupling involving Benzylic and Chemical Sensors. Chem. Rev. 2012, 112, 1105−1125.
(43) Haldar, R.; Bhattacharyya, S.; Maji, T. K. Luminescent Metal-
Allylic C-H Bonds. Org. Chem. Front. 2020, 7, 1717−1742.
Organic Frameworks and Their Potential Applications. J. Chem. Sci.
(24) Gandhi, S. Catalytic Enantioselective Cross Dehydrogenative
2020, 132, 99.
Coupling of sp3 C-H of Heterocycles. Org. Biomol. Chem. 2019, 17,
(44) Zhang, X.; Chen, Z.; Liu, X.; Hanna, S. L.; Wang, X.; Taheri-
9683−9692.
Ledari, R.; Maleki, A.; Li, P.; Farha, O. K. A historical overview of the
(25) He, Z.; Wu, D.; Vessally, E. Cross-dehydrogenative Coupling
activation and porosity of metal-organic frameworks. Chem. Soc. Rev.
Reactions Between Formamidic C(sp2)-H and X-H (X = C, O, N)
2020, 49, 7406−7427.
Bonds. Top. Curr. Chem. 2020, 378, 46. (45) Ji, Z.; Wang, H.; Canossa, S.; Wuttke, S.; Yaghi, O. M. Pore
(26) Yi, H.; Zhang, G.; Wang, H.; Huang, Z.; Wang, J.; Singh, A. K.;
Chemistry of Metal-Organic Frameworks. Adv. Funct. Mater. 2020, 30,
Lei, A. Recent Advances in Radical C−H Activation/Radical Cross- 2000238.
Coupling. Chem. Rev. 2017, 117, 9016−9085. (46) Deria, P.; Mondloch, J. E.; Karagiaridi, O.; Bury, W.; Hupp, J.
(27) Naya, A.; Ishikawa, M.; Matsuda, K.; Ohwaki, K.; Saeki, T.; T.; Farha, O. K. Beyond Post-Synthesis Modification: Evolution of
Noguchi, K.; Ohtake, N. Structure−Activity Relationships of Metal-Organic Frameworks via Building Block Replacement. Chem.
Xanthene Carboxamides, Novel CCR1 Receptor Antagonists. Bioorg. Soc. Rev. 2014, 43, 5896−5912.
Med. Chem. 2003, 11, 875−884. (47) Chen, Y.-Z.; Zhang, R.; Jiao, L.; Jiang, H.-L. Metal-Organic
(28) Lee, S. H.; Nam, D. H.; Park, C. B. Screening Xanthene Dyes Framework-Derived Porous Materials for Catalysis. Coord. Chem. Rev.
for Visible Light-Driven Nicotinamide Adenine Dinucleotide 2018, 362, 1−23.
Regeneration and Photoenzymatic Synthesis. Adv. Synth. Catal. (48) Bavykina, A.; Kolobov, N.; Khan, I. S.; Bau, J. A.; Ramirez, A.;
2009, 351, 2589−2594. Gascon, J. Metal-Organic Frameworks in Heterogeneous Catalysis:
(29) Kim, D.; Park, B.-M.; Yun, J. Highly Efficient Conjugate Recent Progress, New Trends, and Future Perspectives. Chem. Rev.
Reduction of α,β-Unsaturated Nitriles Catalyzed by Copper/ 2020, 120, 8468−8535.
Xanthene-Type Bisphosphine Complexes. Chem. Commun. 2005, (49) Gong, W.; Liu, Y.; Li, H.; Cui, Y. Metal-organic frameworks as
1755−1757. solid Brønsted acid catalysts for advanced organic transformations.
(30) Nishino, H.; Kamachi, H.; Baba, H.; Kurosawa, K. Manganese- Coord. Chem. Rev. 2020, 420, 213400.
(III)-Mediated Carbon-Carbon Bond Formation in the Reaction of (50) Dhakshinamoorthy, A.; Asiri, A. M.; Garcia, H. Catalysis in
Xanthenes with Active Methylene Compounds. J. Org. Chem. 1992, Confined Spaces of Metal Organic Frameworks. ChemCatChem 2020,
57, 3551−3557. 12, 4732−4753.
(31) Piscopo, C. G.; Bühler, S.; Sartori, G.; Maggi, R. Supported (51) Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.;
Sulfonic Acids: Reusable Catalysts for More Sustainable Oxidative Hupp, J. T. Metal-Organic Framework Materials as Catalysts. Chem.
Coupling of Xanthene-Like Compounds with Nucleophiles. Catal. Sci. Soc. Rev. 2009, 38, 1450−1459.
Technol. 2012, 2, 2449−2452. (52) Férey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour,
(32) Zhang, B.; Xiang, S.-K.; Zhang, L. H.; Cui, Y.; Jiao, N. J.; Surblé, S.; Margiolaki, I. A Chromium Terephthalate-Based Solid
Organocatalytic Asymmetric Intermolecular Dehydrogenative alpha- with Unusually Large Pore Volumes and Surface Area. Science 2005,
Alkylation of Aldehydes Using Molecular Oxygen as Oxidant. Org. 309, 2040−2042.
Lett. 2011, 13, 5212−5215. (53) Chen, Y.-Z.; Zhou, Y.-X.; Wang, H.; Lu, J.; Uchida, T.; Xu, Q.;
(33) Chen, Q.; Wang, X.; Yu, G.; Wen, C.; Huo, Y. DDQ-Mediated Yu, S.-H.; Jiang, H.-L. Multifunctional PdAg@MIL-101 for One-Pot
Direct C(sp3)-H Phosphorylation of Xanthene Derivatives. Org. Cascade Reactions: Combination of Host-Guest Cooperation and
Chem. Front. 2018, 5, 2652−2656. Bimetallic Synergy in Catalysis. ACS Catal. 2015, 5, 2062−2069.

10853 https://dx.doi.org/10.1021/acsami.0c20369
ACS Appl. Mater. Interfaces 2021, 13, 10845−10854
ACS Applied Materials & Interfaces www.acsami.org Research Article

(54) Zhao, M.; Yuan, K.; Wang, Y.; Li, G.; Guo, J.; Gu, L.; Hu, W.;
Zhao, H.; Tang, Z. Metal-Organic Frameworks as Selectivity
Regulators for Hydrogenation Reactions. Nature 2016, 539, 76−80.
(55) Henschel, A.; Gedrich, K.; Kraehnert, R.; Kaskel, S. Catalytic
Properties of MIL-101. Chem. Commun. 2008, 4192−4194.
(56) Zhang, Z.; Chen, J.; Bao, Z.; Chang, G.; Xing, H.; Ren, Q.
Insight into the Catalytic Properties and Applications of Metal-
Organic Frameworks in the Cyanosilylation of Aldehydes. RSC Adv.
2015, 5, 79355−79360.
(57) Mortazavi, S.-S.; Abbasi, A.; Masteri-Farahani, M. Influence of
SO3H Groups Incorporated as Brønsted Acidic Parts by Tandem
Post-Synthetic Functionalization on the Catalytic Behavior of MIL-
101(Cr) MOF for Methanolysis of Styrene Oxide. Colloids Surf., A
2020, 599, 124703.
(58) Oudi, S.; Oveisi, A. R.; Daliran, S.; Khajeh, M.; Teymoori, E.
Brønsted-Lewis Dual Acid Sites in a Chromium-Based Metal-Organic
Framework for Cooperative Catalysis: Highly Efficient Synthesis of
Quinazolin-(4H)-1-one Derivatives. J. Colloid Interface Sci. 2020, 561,
782−792.
(59) Mortazavi, S. S.; Abbasi, A.; Masteri-Farahani, M.; Farzaneh, F.
Sulfonic Acid Functionalized MIL-101(Cr) Metal-Organic Frame-
work for Catalytic Production of Acetals. ChemistrySelect 2019, 4,
7495−7501.
(60) Wee, L. H.; Bonino, F.; Lamberti, C.; Bordiga, S.; Martens, J. A.
Cr-MIL-101 encapsulated Keggin Phosphotungstic Acid as Active
Nanomaterial for Catalysing the Alcoholysis of Styrene Oxide. Green
Chem. 2014, 16, 1351−1357.
(61) Santiago-Portillo, A.; Navalón, S.; Cirujano, F. G.; Xamena, F.
X. L. i.; Alvaro, M.; Garcia, H. MIL-101 as Reusable Solid Catalyst for
Autoxidation of Benzylic Hydrocarbons in the Absence of Additional
Oxidizing Reagents. ACS Catal. 2015, 5, 3216−3224.
(62) Chen, X.; Chen, J.; Bao, Z.; Yang, Q.; Yang, Y.; Ren, Q.; Zhang,
Z. MIL-101(Cr)-SO3H Catalyzed Transfer Hydrogenation of 2-
Substituted Quinoline Derivatives. Chin. J. Org. Chem. 2019, 39,
1681−1687.
(63) Chen, J.; Zhang, Z.; Bao, Z.; Su, Y.; Xing, H.; Yang, Q.; Ren, Q.
Functionalized Metal-Organic Framework as a Biomimetic Hetero-
geneous Catalyst for Transfer Hydrogenation of Imines. ACS Appl.
Mater. Interfaces 2017, 9, 9772−9777.
(64) Juan-Alcañiz, J.; Gielisse, R.; Lago, A. B.; Ramos-Fernandez, E.
V.; Serra-Crespo, P.; Devic, T.; Guillou, N.; Serre, C.; Kapteijn, F.;
Gascon, J. Towards Acid MOFs-Catalytic Performance of Sulfonic
Acid Functionalized Architectures. Catal. Sci. Technol. 2013, 3, 2311−
2318.
(65) Akiyama, G.; Matsuda, R.; Sato, H.; Takata, M.; Kitagawa, S.
Cellulose Hydrolysis by a New Porous Coordination Polymer
Decorated with Sulfonic Acid Functional Groups. Adv. Mater. 2011,
23, 3294−3297.
(66) Leveneur, S.; Murzin, D. Y.; Salmi, T. Application of Linear
Free-Energy Relationships to Perhydrolysis of Different Carboxylic
Acids over Homogeneous and Heterogeneous Catalysts. J. Mol. Catal.
A: Chem. 2009, 303, 148−155.
(67) Benfatti, F.; Capdevila, M. G.; Zoli, L.; Benedetto, E.; Cozzi, P.
G. Catalytic Stereoselective Benzylic C−H Functionalizations by
Oxidative C−H Activation and Organocatalysis. Chem. Commun.
2009, 5919−5921.
(68) Banerjee, M.; Das, S.; Yoon, M.; Choi, H. J.; Hyun, M. H.;
Park, S. M.; Seo, G.; Kim, K. Postsynthetic Modification Switches an
Achiral Framework to Catalytically Active Homochiral Metal-Organic
Porous Materials. J. Am. Chem. Soc. 2009, 131, 7524−7525.

10854 https://dx.doi.org/10.1021/acsami.0c20369
ACS Appl. Mater. Interfaces 2021, 13, 10845−10854

You might also like