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

pubs.acs.

org/journal/ascecg Research Article

Heterogeneous Metal Azolate Framework‑6 (MAF-6) Catalysts with


High Zinc Density for Enhanced Polyethylene Terephthalate (PET)
Conversion
Ren-Xuan Yang, Yen-Tsz Bieh, Celine H. Chen, Chang-Yen Hsu, Yuki Kato, Hideki Yamamoto,
Chia-Kuang Tsung,* and Kevin C.-W. Wu*
Cite This: ACS Sustainable Chem. Eng. 2021, 9, 6541−6550 Read Online
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS
Downloaded via CALIFORNIA INST OF TECHNOLOGY on August 11, 2021 at 17:03:26 (UTC).

Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: Polyethylene terephthalate (PET) has been exten-


sively used for the fabrication of various packaging materials,
creating million tons of waste per year. Degrading and recycling
PET waste has been identified as a prominent issue. Herein, we
demonstrate an effective process to chemically convert PET to
bis(2-hydroxyethyl) terephthalate (BHET) through the use of
metal azolate framework-6 (MAF-6) as a catalyst in the presence of
ethylene glycol. MAFs are a subclass of metal−organic frameworks
(MOFs), with MAF-6 comprised of the metal ion Zn2+ and the
organic ligand 2-ethylimidazole. We have optimized the reaction
temperature, reaction time, and catalyst amount to achieve up to a
92.4% conversion of PET and an 81.7% yield of BHET at 180 °C for 4 h. MAF-6 was easily recovered and reused for at least five
times. We have also hypothesized a mechanism for the high conversion and yield of the PET glycolysis reaction catalyzed by MAF-6.
The use of MAF-6 as a catalyst opens a new route for the postconsumer recycling of PET with remarkable practicality.
KEYWORDS: Polyethylene terephthalate (PET), Metal azolate frameworks (MAFs), Glycolysis reaction, Waste recycling treatment,
Polymer depolymerization

■ INTRODUCTION
The global production of synthetic plastics has rapidly grown
into four approaches: namely primary recycling, energy
recovery, mechanical recycling, and chemical recycling.14,15
over the past 50 years due to their diverse range of Among these methods, chemical recycling has attracted
applications, reaching about 350 million metric tons in 2017. considerable attention because the process is less energy
However, approximately 79% of plastic wastes were accumu- intensive and the obtained product can be utilized for the
lated in landfills or the natural environment, 12% were remanufacturing of PET materials.16,17 The chemical recycling
incinerated, and only 9% were recycled.1 Most of the plastic route is a chemical depolymerization process that breaks down
wastes are extremely difficult to degrade in the natural PET ester linkages to form segments and converts the
environment, causing a long-lasting damage to local segments into small-chain fragments. This chemical depolyme-
ecosystems.2−5 Among the wide range of plastic materials, rization can be achieved through hydrolysis,18 methanolysis,19
polyethylene terephthalate (PET), commonly referred to as aminolysis,20 ammonolysis,21 or glycolysis.22 Therefore, PET
“polyester”, has generated an immense amount of waste due to conversion, especially for the chemical depolymerization of
its extensive use in the fabrication of food packaging, soft drink PET, holds great promise for the upcycling of waste PET to
bottles, and fibers in the textile industry.6,7 The robustness and equivalent- or higher-value purposes, such as the synthesis of
durability of PET makes it extremely resistant to chemical new PET or other value-added chemicals.23
degradation. Moreover, the waste PET can be photodegraded Currently, glycolysis is the most promising approach because
into small fragments over time, resulting in microplastics in the main product, bis(2-hydroxyethyl) terephthalate (BHET),
aquatic and marine ecosystems. Eventually, the microplastics
from waste PET can be ingested and accumulated in living Received: November 3, 2020
organisms.8 The annual production of PET exceeded 30 Revised: April 10, 2021
million tons in 2017;9 therefore, researchers have put in Published: May 5, 2021
considerable effort to develop efficient treatments for PET
waste recycling to further convert plastic wastes into useful
materials.10−13 In general, PET recycling can be categorized

© 2021 American Chemical Society https://doi.org/10.1021/acssuschemeng.0c08012


6541 ACS Sustainable Chem. Eng. 2021, 9, 6541−6550
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

can be utilized for the commercial-scale reproduction of the coordinated composition of MAF-6 and ZIF-8 only differs
PET.24 In addition, the reaction occurs under comparatively by one methylene group in the imidazolate ligand, a larger pore
mild reaction conditions with less volatile solvents. Recent size of MAF-6 (16.7 Å) in comparison to ZIF-8 (11.6 Å)48 can
advances in chemical glycolysis reactions have focused on be obtained. In this work, the reaction activity catalyzed by
developing transesterification catalysts to enhance the reaction large-pore MAF-6 and the reaction mechanism are investigated
rate and BHET monomer yield.10 Several efficient catalysts thoroughly.
such as metal oxide,25 nanoparticles,22 ionic liquids (ILs),26 Further developments have examined the activities of an
protic ionic salts,24 and deep-eutectic solvents27−29 have been isomeric series of [Zn(eim)2], including ANA-[Zn(eim)2]
developed. Although they are able to convert PET wastes into (MAF-5) and the nonporous isomer qtz-[Zn(eim)2] (MAF-
BHET monomer with high reaction efficiency under milder 32) with active center similar to that of MAF-6 but with
conditions, some similar demerits of these catalysts still exist different coordination environments.49,50 Here, we report the
such as the requirement of PET pregrinding, low monomer first PET to BHET conversion catalyzed by the large-pore
yields, or elusive separation of products. Consequently, the MAF-6 material, as illustrated in Scheme 1, with the highest
development of new types of catalysts for PET to BHET PET conversion being 92.4% and BHET yield being 81.7%.
conversion is still in high demand.
The study of molecular catalysts has shed light on the design Scheme 1. Schematic Illustration of MAF-6-Catalyzed
of the next-generation catalysts for PET conversion. Early Glycolysis Reaction of PET to BHET
studies have revealed that molecular catalysts show high
activity and monomer yield. For instance, Ghaemy and
Mossaddegh have determined the order of activity of their
catalysts as Zn2+ > Mn2+ > Co2+ > Pb2+.30 It is valuable to
convert these aspects to heterogeneous catalysts, which are
more industrially favorable, because they display easy
separation from reactants and products through filtration and
can bring about less waste production through recycling.26
Metal−organic frameworks (MOFs) provide a solution for
this. MOFs are a new class of crystalline nanoporous materials
self-assembled by metal ions/clusters (joint) and organic
ligands (linkers).31,32 The diverse combinations of metal ions


and organic ligands make MOFs impressive candidates for a
range of applications, including gas storage,33 chemical EXPERIMENTAL SECTION
separation,34 drug delivery,35 electronic devices,36,37 energy
Chemicals. PET pellets (2.0 mm × 2.0 mm × 2.5 mm) were
storage,38 separation,39 and catalysis.40−42 Most importantly, supplied by LIBOLON, Taiwan. Zinc oxide (Sigma-Aldrich, 99%), 2-
the high density and highly tunable metal joints of MOFs ethylimidazole (Sigma-Aldrich, 98%), ammonium hydroxide solution
could serve as active centers and make them promising (Honeywell, ∼25%), ethylene glycol (Alfa Aesar, 99%), methanol
catalysts for PET to BHET conversion. Therefore, they have (Macron, ≥99.8%), and cyclohexane (J.T. Baker, 99.5%) were used as
the potential to display high activity and selectivity as received without further purification. Deionized water was purified
molecular catalysts and are able to be recycled by simple with a Milli-Q system (Millipore, Bedford, MA, USA).
separation from the reaction mixture. The diverse nanoscale Synthesis of Metal Azolate Frameworks (MAFs). MAF-6,
structures and superior chemical stability of zeolitic imidazo- MAF-5, and MAF-32 were synthesized at room temperature by rapid
late frameworks (ZIFs) have received great scientific interest in mixing of an aqueous ammonia solution of zinc oxide (ZnO) and a
methanol solution of 2-ethylimidazole (Heim).50 For MAF-6, the
various applications.43 Indeed, Suo et al. have demonstrated a concentrated aqueous ammonia solution (25% NH4OH, 40 mL) of
result of using zeolitic imidazolate framework-8 (ZIF-8), ZnO (2 mmol) was added dropwise into a methanol (30 mL) and
[Zn(mim)2] (mim = 2-methylimidazole), for PET glycolysis.44 cyclohexane (6.66% v/v) solution of Heim (4 mmol) with rapid
ZIF-8 exhibited a relatively low BHET yield (68.1%) in stirring at room temperature for 0.5 h. For MAF-5, a methanol (6
comparison with the inexpensive zinc chloride (70.0%)45 and mL) solution of Heim (4 mmol) was rapidly poured into a second
other heterogeneous catalysts.28 Notwithstanding the advances solution of ZnO (2 mmol) dissolved in 25% NH4OH (40 mL) and
in PET conversion, the small pore size of ZIF-8 may inhibit the cyclohexane (3.25%, v/v). The solution was stirred at room
activity of PET conversion, and the reaction mechanism is not temperature for 0.5 h. MAF-32 was synthesized through a procedure
similar to that for MAF-5, except that no cyclohexane was added, and
been fully understood as well. The present work aims at
the mixed solutions were stirred for 2 h. After the synthesis, the MAF
establishing a protocol of PET to BHET conversion catalyzed powders were collected by centrifugation and washing with methanol,
by a large-pore MOF material and clarifying the overall repeated three times. The washed powders were stored in a
mechanism. On the basis of the knowledge from molecular lyophilizer overnight for drying.
catalyst studies, we hypothesized that using MOFs with similar General Catalytic Glycolysis Process of PET. In a typical
metal centers while the coordination environment is fine-tuned catalytic test, 5.0 g of PET pellets and 30 g of ethylene glycol (EG)
could promote BHET yield. were placed in a 50 mL round-bottomed two-necked flask equipped
As far as we know, to date, no large-pore MOF catalyst has with a reflux condenser, a thermometer, and a magnetic stirrer.
been applied to PET conversion. To demonstrate that large- Catalysts were placed in the flask, and the flask was immersed in an oil
bath at the target temperature for a predetermined time. The
pore MOF materials, tuned by adjusting the length of the glycolysis reactions were carried out at reaction temperatures ranging
ligand, may render higher reaction efficiency, we selected metal from 160 to 190 °C for 2−6 h under atmospheric pressure. When
azolate framework-6 (MAF-6), RHO-[Zn(eim)2], which is each reaction was complete, the reaction mixture was cooled to room
comprised of Zn2+ ions coordinated by 2-ethylimidazole temperature and an excess amount of distilled water (300−400 mL)
(Heim) linkers to use in the PET conversion.46,47 Although was added to the solution with vigorous stirring. The catalysts and

6542 https://doi.org/10.1021/acssuschemeng.0c08012
ACS Sustainable Chem. Eng. 2021, 9, 6541−6550
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 1. MAF-6 powder characterization. (a) PXRD patterns of the as-prepared and simulated MAF-6; (b) SEM image of MAF-6; (c) N2
adsorption (filled dots) and desorption (empty dots) isotherms of MAF-6; (d) TGA curve of MAF-6.

undepolymerized PET were filtered and separated from the products. a nitrogen sorption analysis (BELSORP-max II analyzer). The
The undepolymerized PET was dried at 70 °C for 12 h and weighed Brunauer−Emmett−Teller (BET) method was used to model the
for the calculation of PET conversion, which is defined by eq 1 specific surface area. The crystallinity and morphology of the catalysts
were analyzed by powder X-ray diffraction (PXRD, Rigaku) and
Wi − Wf
PET conversion (%) = × 100% scanning electron microscopy (SEM, NovaTM NanoSEM 230). The
Wi (1) strength and type of acidic sites on the MAFs were evaluated by the
temperature-programmed desorption of adsorbed ammonia (NH3-
where Wi represents the initial weight of PET and Wf represents the TPD) and IR spectra of adsorbed pyridine (see Figure S3 in the
weight of undepolymerized PET. In order to separate the oligomers Supporting Information). The reaction products were analyzed by an
and dimers from the residual PET, 500 mL of distilled water was elemental analyzer (EA, Elementar vario EL, cube type), a
added to the residual PET and the solution was boiled and stirred for
thermogravimetric analyzer (TGA/DSC, TA Instruments SDT
45 min. The residue on the filter was an oligomer, and the filtrate was
650), nuclear magnetic resonance (1H NMR, Bruker drx500), and a
immersed in an ice bath. The white floccules that formed were
gas chromatograph−mass spectrometer (GC-MS, Agilent 6890/
considered to be dimers. Moreover, the collected aqueous transparent
5975).
filtrate was concentrated to about 150−200 mL and stored in a
refrigerator at 4 °C overnight. White needle-like crystals were formed
in the solution and then filtered, dried, and weighed. The DSC scan,
GC-MS, and NMR characterization revealed that these crystals are
indeed BHET monomers. The BHET molar yield is defined by eq 2
■ RESULTS AND DISCUSSION
Characterization of As-Synthesized MAF Catalysts.
The morphology of MAF-6 was analyzed by SEM (Figure 1b).
WBHET/MWBHET In the SEM image, MAF-6 displays a homogeneous cubic
BHET yield (%) = × 100%
WPET/MWPET (2) structure with particle sizes of 250−300 nm, and MAF-32 is
constructed by block-shaped particles with sizes of 200−300
where WBHET and WPET correspond to the initial weight of PET and
nm. MAF-5 has a nonhomogeneous polyhedral structure with
the weight of obtained BHET, respectively. MWBHET and MWPET
refer to the molar mass of BHET (254 g mol−1) and the PET larger sizes of 3−15 μm (see Figure S1 in the Supporting
repeating unit (192 g mol−1), respectively. Information). The PXRD patterns of MAF-6 fit well with
Characterization. The surface area, porosity, pore size, and pore simulated patterns, which confirms the formation of MAF-6,
volume of the as-synthesized MAF catalysts were determined by using and show no other crystalline impurity phases (Figure 1a). In
6543 https://doi.org/10.1021/acssuschemeng.0c08012
ACS Sustainable Chem. Eng. 2021, 9, 6541−6550
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 2. (a) Effect of amount of catalyst on PET glycolysis (temperature 180 °C, time 4 h). (b) Effect of temperature on PET glycolysis (catalyst
amount 0.05 g, time 4 h). (c) Effect of time on PET glycolysis (catalyst amount 0.05 g, temperature 180 °C). (d) Reusability of MAF-6 in PET
glycolysis (catalyst amount 0.05 g, temperature 180 °C, time 4 h).

Table 1. PET Glycolysis Catalyzed by Different Catalystsa Table 2. Hansen Solubility Parameters of Specific Solvents
Used for Solubility Tests
catalyst PET conversion (%) BHET yield (%)
no catalyst 0 0 entry solvent δd (MPa1/2) δp (MPa1/2) δh (MPa1/2)
zinc acetateb 71.0 51.8 1 hexane 14.9 0 0
zinc chlorideb 58.0 34.8 2 toluene 18 1.4 2
zinc oxideb 28.2 9.0 3 nitrosobenzene 20 12.7 4
MAF-5b 72.3 39.0 4 methyl ethyl ketone 16 9 5.1
MAF-6b 92.4 81.7 5 quinoline 20.5 5.6 5.7
MAF-32b 52.6 38.2 6 acetone 15.5 10.4 7
a 7 ethyl acetate 15.8 5.3 7.2
Reaction conditions: PET (5.0 g), EG (30 g), atmospheric pressure
at 180 °C, 4 h. bZinc acetate, zinc chloride, zinc oxide, MAF-5, MAF- 8 1,4-dioxane 17.5 1.8 9
6, and MAF-32 were added to the PET glycolysis system with the 9 dimethyl sulfoxide 18.4 16.4 10.2
same percentage content of Zn based on MAF-6 (7.7 mmol). The 10 dimethylformamide 17.4 13.7 11.3
actual amount of Zn was measured by ICP (see Table S2 in the 11 benzyl alcohol 18.4 6.3 13.7
Supporting Information). 12 N-methylformamide 17.4 18.8 15.9
13 1-propanol 16 6.8 17.4
addition, the strong intensities of the diffraction peaks suggest 14 diethylene glycol 16.6 12 19
that MAF-6 possesses high crystallinity. The structures of 15 formamide 17.2 26.2 19
MAF-5 and MAF-32 were confirmed by PXRD as well (see 16 ethanol 15.8 8.8 19.4
Figure S2a in the Supporting Information). N2 adsorption− 17 ethanolamine 17 15.5 21
desorption isotherms show a high Brunauer−Emmett−Teller 18 methanol 14.7 12.3 22.3
(BET) specific surface area of the MAF-6 sample (Figure 1c) 19 ethylene glycol 17 11 26
of approximately 1544 m2/g and a pore size of around 16.7 Å
(see Table S1 in the Supporting Information). MAF-6 exhibits MAF-6 was analyzed by thermogravimetric analysis (TGA)
a reversible type I physisorption isotherm on the basis of the under the N2 atmosphere (Figure 1d). The TGA result of
characterization of the International Union of Pure and MAF-6 showed no significant weight loss below 400 °C,
Applied Chemistry (IUPAC) classification, with a sharp indicating the good thermal stability of MAF-6. The thermal
increase at a low relative pressure, suggesting that MAF-6 stabilities of MAF-5 and MAF-32 were also analyzed by TGA
possesses a microporous feature. The N2 physisorption (see Figure S2c in the Supporting Information).
isotherms of MAF-5 and MAF-32 are displayed in Figure Catalytic Glycolysis Reaction Optimization of PET.
S2b in the Supporting Information. The thermal stability of Due to the high performance of MAF-6, we studied the
6544 https://doi.org/10.1021/acssuschemeng.0c08012
ACS Sustainable Chem. Eng. 2021, 9, 6541−6550
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Table 3. Solubility Scores of MOF Materials in Certain Organic Solventsa


entry solvent MAF-5 MAF-6 MAF-32 ZIF-8 ZIF-67
1 hexane 1 0 0 0 0
2 toluene 2 0 0 0 0
3 nitrobenzene 0 0 0 0 0
4 methyl ethyl ketone 0 0 0 0 0
5 quinoline 0 0 0 0 1
6 acetone 0 0 0 1 0
7 ethyl acetate 0 0 0 1 0
8 1,4-dioxane 0 0 0 0 0
9 dimethyl sulfoxide 0 0 0 0 0
10 dimethylformamide 0 0 0 0 0
11 benzyl alcohol 0 0 0 0 0
12 N-methylformamide 0 1 0 0 1
13 1-propanol 0 1 0 0 0
14 diethylene glycol 0 0 1 0 1
15 formamide 0 1 0 0 0
16 ethanol 0 0 0 0 0
17 ethanolamine 0 0 0 0 0
18 methanol 0 0 0 1 0
19 ethylene glycol 0 0 1 0 1
a
The value of the score was determined by 1 and 2 as being a good solvent and 0 as being a poor solvent.

Figure 3. Three-dimensional HSP plots and Hansen solubility spheres for each target compound. Blue spheres represent “good” solvents, and red
cubes represent “poor” solvents.

catalysis of MAF-6 in detail. The PET glycolysis over the 92.4% and 81.7%, respectively. The reason may be that the
MAF-6 catalyst was carried with different amounts of catalyst, smaller amount of catalyst possessed less surface area and
reaction temperatures, and reaction times (Figure 2). First, the fewer active sites in comparison to that of 0.05 g of catalyst,
PET glycolysis was performed at 180 °C for 4 h with variation decelerating the process of reaching a chemical equilibrium.
of the amounts of MAF-6 catalyst (Figure 2a). When the However, when the amount of catalyst was increased further to
amount of catalyst was 0.005 g, the conversion of PET and 0.5 g, the conversion of PET and yield of BHET decreased to
yield of BHET were 58.6% and 48.8%, respectively. By an 89.6% and 60.3%, respectively. The possible reason is that
increase in the amount of catalyst added to the glycolysis monomers will undergo a polymerization reaction by
system, the conversion of PET and yield of BHET showed a themselves and occupy some active sites, which may cause a
significant increase. When the amount of catalyst was raised to slowing of the reaction rate and obtaining a lower yield of
0.05 g, the conversion of PET and yield of BHET reached BHET. Thus, the catalyst amount of 0.05 g was chosen to be
6545 https://doi.org/10.1021/acssuschemeng.0c08012
ACS Sustainable Chem. Eng. 2021, 9, 6541−6550
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

both the conversion and yield increased with an increase in


time. When the reaction time was extended to 4 h, 92.4%
conversion of PET could be achieved at 180 °C. Nevertheless,
if the reaction time was prolonged to 6 h, the yield of BHET
decreased, which was caused by the chemical reaction
equilibrium between the BHET monomer and the dimer. In
short, with a reaction time extending to 6 h, the dimerization of
BHET was carried out as a side reaction. Thus, it can be
concluded that the operating time is a critical factor for the
glycolysis of PET.
The feasibility of recycling this heterogeneous catalyst was
also evaluated in this study, as shown in Figure 2d. At the end
of each run, the catalyst was separated from the reaction
mixture by filtration and directly reused for a subsequent
reaction under the optimized reaction conditions, which were
180 °C for 4 h under atmospheric pressure. After five runs, the
catalytic performance retained a 71.2% product yield. From the
XRD pattern, shown in Figure S3c in the Supporting
Information, the phase composition of used MAF-6 is same
as that of fresh MAF-6. In addition, an ICP-MS measurement
Figure 4. Three-dimensional HSP of target materials and the affinity revealed that only 3.7% of metal ions leached out after five
(Ra) between PET and samples in ethylene glycol.
times of recycling. These results demonstrate that the high
stability of MAF-6 provides a sustainable and practical future
Table 4. Hansen Solubility Parameters of Specific
application for the PET glycolysis reaction.
Substances in the 3D Diagram (Figure 4)
In the catalytic evaluation tests, the reactions were catalyzed
by as-synthesized MAFs and other zinc salt catalysts, and the
results are summarized in Table 1. In the catalysis by MAF-6,
PET is depolymerized with a particularly higher yield of 81.7%
BHET in comparison with conventional metal salts reported in
previous literature. The reason can be ascribed to the high
specific surface area of MAF-6. Thus, aside from this, the
solubility discrepancies of the reactant and catalyst in ethylene
glycol may affect the yield of BHET, as an important factor in
this system.
Affinity Evaluation. The solvent usually plays a crucial
role in the uniformity of a reaction mixture. The solubilities of
the reactant and catalyst in organic solvents are important
properties for giving high catalytic performance in both
conversion and yield. The Hansen solubility parameters
(HSPs) have recently attracted attention and are widely used
to evaluate the solubility of materials. The HSPs are comprised
of δd, δp, and δh (MPa1/2), which are dispersion, polar, and
hydrogen-bonding forces in the cohesive energy density of the
target compound, respectively.51 In addition, the solubility
parameter (δt) can be expressed as three partial terms, as
applied in the other reactions because of the highest shown in eq 3:
conversion of PET and yield of BHET. The characterizations
of the product are shown in Table S3 and Figures S4−S6 in the δt = (δd 2 + δp2 + δ h 2)1/2 (3)
Supporting Information. Figure 2b presents the influence of
reaction temperature on PET conversion and BHET yield. The The HSPs of MOF catalysts were determined via the Hansen
conversion of PET increased significantly with an increase in solubility sphere method for an affinity evaluation. The series
temperature from 160 to 180 °C, with the maximum value of 19 solvents with a wide range of solubility parameters shown
reaching 92.4% at 180 °C with an 81.7% BHET yield. The in Table 2 were selected for solubility tests. The corresponding
results revealed that high temperature could afford sufficient dispersion stability scores for the target material in each
energy to expedite the catalytic reaction rate, and the PET organic solvent are shown in Table 3. An example of the score
polymer chain could be broken down into monomers. measured experimentally by the solubility test, known as a
However, the PET conversion (94.0%) and BHET yield dispersion stability evaluation in certain types of solvents, is
(78.3%) decreased slightly when the temperature was further shown in Figure S7 in the Supporting Information. Figure 3
increased to 190 °C, indicating that a reaction temperature displays the three-dimensional (3D) HSP plots of good and
over 180 °C would affect the reaction equilibrium between the poor solvents for each target substance on the basis of the
BHET monomer and the dimer.27 solubility test results. Blue spheres and red cubes represent
A time-dependent analysis of the PET glycolysis catalyzed solvents with good and poor affinity, respectively. When the
by MAF-6 is presented in Figure 2c. In the period of 1−4 h, scores obtained from solubility tests are input into the
6546 https://doi.org/10.1021/acssuschemeng.0c08012
ACS Sustainable Chem. Eng. 2021, 9, 6541−6550
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Scheme 2. Proposed Reaction Mechanism for PET Glycolysis Catalyzed by MAF-6

Scheme 3. Size of MAF-6 and of the Dimer and Monomer (BHET)

calculation program, Hansen solubility parameters in practice Ra =


(HSPiP) made by Steven Abbott TCNF Ltd. UK (5th
[4 × (δd,1 − δd,2)2 + (δp,1 − δp,2)2 + (δ h,1 − δ h,2)2 ]1/2
edition),52 show that a good solvent in a certain region can
(4)
form a sphere in the 3D diagram representing the Hansen
solubility sphere. Good solvents are plotted inside the where δd,1, δp,1, and δh,1 represent the HSP values of
component 1 and δd,2, δp,2, and δh,2 represent the HSP values
solubility sphere, while poor solvents are plotted outside. of component 2. The HSPs of each material can be illustrated
Therefore, the central coordinate of the solubility sphere can in a 3D diagram, as shown in Figure 4. A smaller Ra value
be defined as the HSP value of the target substance. means a higher affinity of each substance because the
The index described as Ra in eq 4 represents the distance interaction forces between the molecules are similar. Table 4
shows the HSPs of specific substances and the distance
between the HSPs of two substances in the 3D diagram of the
between the HSPs of MOF materials from ethylene glycol
Hansen sphere and can be used to evaluate the affinity between (EG) and PET. The affinities of the PET reactant and MOF
substances quantitatively catalyst in ethylene glycol are critical values in achieving good
6547 https://doi.org/10.1021/acssuschemeng.0c08012
ACS Sustainable Chem. Eng. 2021, 9, 6541−6550
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 5. Reactions with different PET feedstocks.

catalytic performance in the PET glycolysis system. Therefore, the reaction rate can be accelerated due to the abundant active
both values should be taken into consideration. A better sites in the pore structure. The larger pore size of MAF-6
catalytic performance can be attributed to the case of the exhibits a selective diffusion of dimers into pores for further
similar interaction distances (Ra) of the MOF catalyst from EG conversion into BHET monomers. Chen et al. found a similar
and PET. Among all the MOF catalysts included in the HSPs result in the selective hydrogenation of cinnamaldehyde that
test, MAF-6 has the most similar affinity in Ra index, indicating the flexible framework of ZIF-8 can allow cinnamaldehyde to
great compatibility and good dispersibility between the PET pass more easily through the pores and contact the active sites
reactant and MAF-6 catalyst in the organic solvent, rendering in the pore structure in comparison to the rigid framework of
higher possibilities of effective collisions and interactions ZIF-71.55
between two particles during the PET glycolysis reaction and The reaction protocol is also applicable in the catalysis of
further enhancing the yield of BHET. postconsumer PET bottles. As shown in Figure 5, on
Reaction Mechanism of PET Catalytic Glycolysis. The conversion of wasted PET beverage bottles into BHET by
reaction mechanism of PET glycolysis catalyzed by MAF-6 is means of a glycolysis reaction, a 92% conversion of PET and
proposed in Scheme 2. Zinc ions of MAF-6 act as a Lewis acid 61.7% yield of BHET can be achieved. Thus, the reaction
to catalyze consecutive transesterification reactions.25,27,29,53,54 strategy in this work is considered to be a promising route for
These are SN2 nucleophilic substitution reactions, in which the depolymerization reactions, rendering greater possibilities in
lone pair electrons of the carboxyl oxygen in the ester linkages further practical applications.
of the PET unit are attacked. In the first step, Zn ions on the
surface of MAF-6 attack the carboxyl oxygen and facilitate
polarizing carboxyl oxygen. This process further enhances the
■ CONCLUSION
In summary, we have developed a catalytic process of PET
electrophilicity of the adjacent carbon. Therefore, the oxygen glycolysis over MAF-6 catalyst in the presence of ethylene
in the hydroxyl of ethylene glycol is capable of nucleophilic glycol. This process achieved an efficient and sustainable PET
attack at the carboxyl carbon, resulting in a tetrahedral recycling. The outstanding catalytic performance can be
transition state, further pushing off the metal ion, and breaking attributed to MAF-6’s high specific surface area, large pores,
the ester linkage of PET. Afterward, the oligomers and dimers and good dispersibility between PET and ethylene glycol. A
are generated sequentially by means of continuously repeating remarkable yield of BHET (81.7%) was achieved in the
these reaction processes. Eventually, the desired monomeric presence of MAF-6 (0.05 g), ethylene glycol (30 g), and PET
product, BHET, can be formed. On the basis of this reaction (5.0 g) for 4 h of glycolysis at 180 °C. The MAF-6 catalyst can
mechanism, we can assume that the Zn node on the external be recycled for at least five sequential runs without sacrificing
surface of MAF-6 can do the first few steps. After more bonds the performance. The results show that the large pore size of
are broken into small fragments, such as dimers, the reaction MAF-6 can provide guidance for the design of more efficient
can be catalyzed by the Zn node in the pore of the MAF-6 and selective catalysts, which may be taken into consideration
scaffold. for continuous conversion of PET wastes by the process
For the sake of verifying the dimers’ abilities to diffuse into design. Consequently, this work is capable of providing a
the pores of MAF-6, the size effect was examined and is highly efficient recycling route of PET wastes and shows
displayed in Scheme 3. The molecular size of dimers with promising prospects for further industrial applications.


minimized energy turns out to be capable of passing through
the large pores and generating BHET. In contrast, the smaller
ASSOCIATED CONTENT
pore size of MAF-5 and nonporous MAF-32 (see Table S1 in
the Supporting Information) made it difficult for dimers to * Supporting Information

diffuse into the pore structure. This result demonstrates that The Supporting Information is available free of charge at
dimers can diffuse into the pore structure of MAF-6, so that https://pubs.acs.org/doi/10.1021/acssuschemeng.0c08012.
6548 https://doi.org/10.1021/acssuschemeng.0c08012
ACS Sustainable Chem. Eng. 2021, 9, 6541−6550
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Characterization of as-synthesized MAFs catalysts, (5) Rochman, C. M. Microplastics research-from sink to source.
characterization of PET glycolysis products, and Hansen Science 2018, 360 (6384), 28−29.
solubility parameters of the catalysts (PDF) (6) Jehanno, C.; Pérez-Madrigal, M. M.; Demarteau, J.; Sardon, H.;


Dove, A. P. Organocatalysis for depolymerisation. Polym. Chem. 2019,
10 (2), 172−186.
AUTHOR INFORMATION (7) Zhou, L.; Lu, X.; Ju, Z.; Liu, B.; Yao, H.; Xu, J.; Zhou, Q.; Hu, Y.;
Corresponding Authors Zhang, S. Alcoholysis of polyethylene terephthalate to produce dioctyl
terephthalate using choline chloride-based deep eutectic solvents as
Chia-Kuang Tsung − Department of Chemistry, Merkert
efficient catalysts. Green Chem. 2019, 21 (4), 897−906.
Chemistry Center, Boston College, Chestnut Hill, (8) Yuan, X.; Lee, J. G.; Yun, H.; Deng, S.; Kim, Y. J.; Lee, J. E.;
Massachusetts 02467, United States; orcid.org/0000- Kwak, S. K.; Lee, K. B. Solving two environmental issues
0002-9410-565X; Email: frank.tsung@bc.edu simultaneously: Waste polyethylene terephthalate plastic bottle-
Kevin C.-W. Wu − Department of Chemical Engineering, derived microporous carbons for capturing CO2. Chem. Eng. J.
National Taiwan University (NTU), Taipei 10617, Taiwan; 2020, 397, 125350.
Center of Atomic Initiative for New Materials (AI-MAT), (9) Danso, D.; Chow, J.; Streit, W. R. Plastics: Environmental and
National Taiwan University, Taipei 10617, Taiwan; Biotechnological Perspectives on Microbial Degradation. Appl.
orcid.org/0000-0003-0590-1396; Email: kevinwu@ Environ. Microbiol. 2019, 85 (19), No. e01095-19.
ntu.edu.tw (10) Al-Sabagh, A. M.; Yehia, F. Z.; Eshaq, G.; Rabie, A. M.;
ElMetwally, A. E. Greener routes for recycling of polyethylene
Authors terephthalate. Egypt. J. Pet. 2016, 25 (1), 53−64.
Ren-Xuan Yang − Department of Chemical Engineering, (11) Fukushima, K.; Coulembier, O.; Lecuyer, J. M.; Almegren, H.
National Taiwan University (NTU), Taipei 10617, Taiwan; A.; Alabdulrahman, A. M.; Alsewailem, F. D.; Mcneil, M. A.; Dubois,
P.; Waymouth, R. M.; Horn, H. W.; Rice, J. E.; Hedrick, J. L.
Center of Atomic Initiative for New Materials (AI-MAT),
Organocatalytic depolymerization of poly(ethylene terephthalate). J.
National Taiwan University, Taipei 10617, Taiwan; Polym. Sci., Part A: Polym. Chem. 2011, 49 (5), 1273−1281.
orcid.org/0000-0001-9589-629X (12) Malik, N.; Kumar, P.; Shrivastava, S.; Ghosh, S. B. An overview
Yen-Tsz Bieh − Department of Chemical Engineering, on PET waste recycling for application in packaging. Int. J. Plast.
National Taiwan University (NTU), Taipei 10617, Taiwan Technol. 2017, 21 (1), 1−24.
Celine H. Chen − School of Engineering, Brown University, (13) Raheem, A. B.; Noor, Z. Z.; Hassan, A.; Abd Hamid, M. K.;
Providence, Rhode Island 02912, United States Samsudin, S. A.; Sabeen, A. H. Current developments in chemical
Chang-Yen Hsu − Department of Chemical Engineering, recycling of post-consumer polyethylene terephthalate wastes for new
National Taiwan University (NTU), Taipei 10617, Taiwan materials production: A review. J. Cleaner Prod. 2019, 225, 1052−
Yuki Kato − Department of Chemical and Environmental 1064.
Engineering, Faculty of Environmental and Urban (14) Al-Salem, S. M.; Lettieri, P.; Baeyens, J. Recycling and recovery
Engineering, Kansai University, Suita, Osaka 564-8680, routes of plastic solid waste (PSW): a review. Waste Manage. 2009, 29
(10), 2625−2643.
Japan; orcid.org/0000-0002-8520-3389 (15) Garcia, J. M.; Robertson, M. L. The future of plastics recycling.
Hideki Yamamoto − Department of Chemical and Science 2017, 358 (6365), 870−872.
Environmental Engineering, Faculty of Environmental and (16) Dyosiba, X.; Ren, J.; Musyoka, N. M.; Langmi, H. W.; Mathe,
Urban Engineering, Kansai University, Suita, Osaka 564- M.; Onyango, M. S. Feasibility of Varied Polyethylene Terephthalate
8680, Japan; orcid.org/0000-0001-7113-0070 Wastes as a Linker Source in Metal-Organic Framework UiO-66(Zr)
Complete contact information is available at: Synthesis. Ind. Eng. Chem. Res. 2019, 58 (36), 17010−17016.
(17) Sinha, V.; Patel, M. R.; Patel, J. V. Pet Waste Management by
https://pubs.acs.org/10.1021/acssuschemeng.0c08012
Chemical Recycling: A Review. J. Polym. Environ. 2010, 18 (1), 8−25.
(18) Paliwal, N. R.; Mungray, A. K. Ultrasound assisted alkaline
Notes hydrolysis of poly(ethylene terephthalate) in presence of phase
The authors declare no competing financial interest.


transfer catalyst. Polym. Degrad. Stab. 2013, 98 (10), 2094−2101.
(19) Kurokawa, H.; Ohshima, M.-a.; Sugiyama, K.; Miura, H.
ACKNOWLEDGMENTS Methanolysis of polyethylene terephthalate (PET) in the presence of
This work was supported by the Ministry of Science and aluminium tiisopropoxide catalyst to form dimethyl terephthalate and
Technology (MOST), Taiwan under the Shackleton Program ethylene glycol. Polym. Degrad. Stab. 2003, 79 (3), 529−533.
(20) Tawfik, M. E.; Eskander, S. B. Chemical recycling of
award number 108-2638-E-002-003-MY2, and the other poly(ethylene terephthalate) waste using ethanolamine. Sorting of
MOST grant (109-2917-I-564-015).


the end products. Polym. Degrad. Stab. 2010, 95 (2), 187−194.
(21) Mittal, A.; Soni, R. K.; Dutt, K.; Singh, S. Scanning electron
REFERENCES microscopic study of hazardous waste flakes of polyethylene
(1) Geyer, R.; Jambeck, J. R.; Law, K. L. Production, use, and fate of terephthalate (PET) by aminolysis and ammonolysis. J. Hazard.
all plastics ever made. Sci. Adv. 2017, 3 (7), No. e1700782. Mater. 2010, 178 (1−3), 390−396.
(2) Jambeck, J. R.; Geyer, R.; Wilcox, C.; Siegler, T. R.; Perryman, (22) Veregue, F. R.; Pereira da Silva, C. T.; Moisés, M. P.;
M.; Andrady, A.; Narayan, R.; Law, K. L. Marine pollution. Plastic Meneguin, J. G.; Guilherme, M. R.; Arroyo, P. A.; Favaro, S. L.;
waste inputs from land into the ocean. Science 2015, 347 (6223), Radovanovic, E.; Girotto, E. M.; Rinaldi, A. W. Ultrasmall Cobalt
768−771. Nanoparticles as a Catalyst for PET Glycolysis: A Green Protocol for
(3) Lamb, J. B.; Willis, B. L.; Fiorenza, E. A.; Couch, C. S.; Howard, Pure Hydroxyethyl Terephthalate Precipitation without Water. ACS
R.; Rader, D. N.; True, J. D.; Kelly, L. A.; Ahmad, A.; Jompa, J.; Sustainable Chem. Eng. 2018, 6 (9), 12017−12024.
Harvell, C. D. Plastic waste associated with disease on coral reefs. (23) Kosloski-Oh, S. C.; Wood, Z. A.; Manjarrez, Y.; de los Rios, J.
Science 2018, 359 (6374), 460−462. P.; Fieser, M. E. Catalytic methods for chemical recycling or upcycling
(4) Lee, H.; Kunz, A.; Shim, W. J.; Walther, B. A. Microplastic of commercial polymers. Mater. Horiz. 2021, 8, 1084.
contamination of table salts from Taiwan, including a global review. (24) Jehanno, C.; Flores, I.; Dove, A. P.; Müller, A. J.; Ruipérez, F.;
Sci. Rep. 2019, 9 (1), 10145. Sardon, H. Organocatalysed depolymerisation of PET in a fully

6549 https://doi.org/10.1021/acssuschemeng.0c08012
ACS Sustainable Chem. Eng. 2021, 9, 6541−6550
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

sustainable cycle using thermally stable protic ionic salt. Green Chem. (42) Nguyen, C. V.; Liao, Y.-T.; Kang, T.-C.; Chen, J. E.; Yoshikawa,
2018, 20 (6), 1205−1212. T.; Nakasaka, Y.; Masuda, T.; Wu, K. C. W. A metal-free, high
(25) Imran, M.; Kim, D. H.; Al-Masry, W. A.; Mahmood, A.; nitrogen-doped nanoporous graphitic carbon catalyst for an effective
Hassan, A.; Haider, S.; Ramay, S. M. Manganese-, cobalt-, and zinc- aerobic HMF-to-FDCA conversion. Green Chem. 2016, 18 (22),
based mixed-oxide spinels as novel catalysts for the chemical recycling 5957−5961.
of poly(ethylene terephthalate) via glycolysis. Polym. Degrad. Stab. (43) Kaneti, Y. V.; Dutta, S.; Hossain, M. S. A.; Shiddiky, M. J. A.;
2013, 98 (4), 904−915. Tung, K.-L.; Shieh, F.-K.; Tsung, C.-K.; Wu, K. C.-W.; Yamauchi, Y.
(26) Wang, Q.; Geng, Y.; Lu, X.; Zhang, S. First-Row Transition Strategies for Improving the Functionality of Zeolitic Imidazolate
Metal-Containing Ionic Liquids as Highly Active Catalysts for the Frameworks: Tailoring Nanoarchitectures for Functional Applica-
Glycolysis of Poly(ethylene terephthalate) (PET). ACS Sustainable tions. Adv. Mater. 2017, 29 (38), 1700213.
Chem. Eng. 2015, 3 (2), 340−348. (44) Suo, Q.; Zi, J.; Bai, Z.; Qi, S. The Glycolysis of Poly(ethylene
(27) Liu, B.; Fu, W.; Lu, X.; Zhou, Q.; Zhang, S. Lewis Acid-Base terephthalate) Promoted by Metal Organic Framework (MOF)
Synergistic Catalysis for Polyethylene Terephthalate Degradation by Catalysts. Catal. Lett. 2017, 147 (1), 240−252.
1,3-Dimethylurea/Zn(OAc)2 Deep Eutectic Solvent. ACS Sustainable (45) López-Fonseca, R.; Duque-Ingunza, I.; de Rivas, B.; Arnaiz, S.;
Chem. Eng. 2019, 7 (3), 3292−3300. Gutiérrez-Ortiz, J. I. Chemical recycling of post-consumer PET wastes
(28) Liu, Y.; Yao, X.; Yao, H.; Zhou, Q.; Xin, J.; Lu, X.; Zhang, S. by glycolysis in the presence of metal salts. Polym. Degrad. Stab. 2010,
Degradation of poly(ethylene terephthalate) catalyzed by metal-free 95 (6), 1022−1028.
choline-based ionic liquids. Green Chem. 2020, 22 (10), 3122−3131. (46) Zhang, J. P.; Zhang, Y. B.; Lin, J. B.; Chen, X. M. Metal azolate
(29) Wang, Q.; Yao, X.; Geng, Y.; Zhou, Q.; Lu, X.; Zhang, S. Deep frameworks: from crystal engineering to functional materials. Chem.
eutectic solvents as highly active catalysts for the fast and mild Rev. 2012, 112 (2), 1001−1033.
(47) Zhang, X.-W.; Jiang, L.; Mo, Z.-W.; Zhou, H.-L.; Liao, P.-Q.;
glycolysis of poly(ethylene terephthalate)(PET). Green Chem. 2015,
Ye, J.-W.; Zhou, D.-D.; Zhang, J.-P. Nitrogen-doped porous carbons
17 (4), 2473−2479.
derived from isomeric metal azolate frameworks. J. Mater. Chem. A
(30) Ghaemy, M.; Mossaddegh, K. Depolymerisation of poly-
2017, 5 (46), 24263−24268.
(ethylene terephthalate) fibre wastes using ethylene glycol. Polym.
(48) Park, K. S.; Ni, Z.; Cote, A. P.; Choi, J. Y.; Huang, R.; Uribe-
Degrad. Stab. 2005, 90 (3), 570−576. Romo, F. J.; Chae, H. K.; O’Keeffe, M.; Yaghi, O. M. Exceptional
(31) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. The chemical and thermal stability of zeolitic imidazolate frameworks.
chemistry and applications of metal-organic frameworks. Science 2013, Proc. Natl. Acad. Sci. U. S. A. 2006, 103 (27), 10186−10191.
341 (6149), 1230444−1230456. (49) Beldon, P. J.; Fabian, L.; Stein, R. S.; Thirumurugan, A.;
(32) Yaghi, O. M.; O’Keeffe, M.; Ockwig, N. W.; Chae, H. K.; Cheetham, A. K.; Friscic, T. Rapid room-temperature synthesis of
Eddaoudi, M.; Kim, J. Reticular synthesis and the design of new zeolitic imidazolate frameworks by using mechanochemistry. Angew.
materials. Nature 2003, 423 (6941), 705−714. Chem., Int. Ed. 2010, 49 (50), 9640−9643.
(33) Peng, Y.; Krungleviciute, V.; Eryazici, I.; Hupp, J. T.; Farha, O. (50) He, C. T.; Jiang, L.; Ye, Z. M.; Krishna, R.; Zhong, Z. S.; Liao,
K.; Yildirim, T. Methane storage in metal-organic frameworks: current P. Q.; Xu, J.; Ouyang, G.; Zhang, J. P.; Chen, X. M. Exceptional
records, surprise findings, and challenges. J. Am. Chem. Soc. 2013, 135 Hydrophobicity of a Large-Pore Metal-Organic Zeolite. J. Am. Chem.
(32), 11887−11894. Soc. 2015, 137 (22), 7217−7223.
(34) Herm, Z. R.; Wiers, B. M.; Mason, J. A.; van Baten, J. M.; (51) Hansen, C. M. Hansen Solubility Parameters: A User’s
Hudson, M. R.; Zajdel, P.; Brown, C. M.; Masciocchi, N.; Krishna, R.; Handbook, 2nd ed.; CRC Press: 1999.
Long, J. R. Separation of hexane isomers in a metal-organic framework (52) Abbott, S.; Hansen, C. M.; Yamamoto, H. S.; Valpey, R. S.
with triangular channels. Science 2013, 340 (6135), 960−964. Hansen Solubility Parameters in Practice Complete with eBook., 5th ed.;
(35) Horcajada, P.; Chalati, T.; Serre, C.; Gillet, B.; Sebrie, C.; Baati, 2015.
T.; Eubank, J. F.; Heurtaux, D.; Clayette, P.; Kreuz, C.; Chang, J. S.; (53) Apicella, B.; Di Serio, M.; Fiocca, L.; Po, R.; Santacesaria, E.
Hwang, Y. K.; Marsaud, V.; Bories, P. N.; Cynober, L.; Gil, S.; Ferey, Kinetic and catalytic aspects of the formation of poly(ethylene
G.; Couvreur, P.; Gref, R. Porous metal-organic-framework nanoscale terephthalate) (PET) investigated with model molecules. J. Appl.
carriers as a potential platform for drug delivery and imaging. Nat. Polym. Sci. 1998, 69 (12), 2423−2433.
Mater. 2010, 9 (2), 172−178. (54) Reinoso, D. M.; Ferreira, M. L.; Tonetto, G. M. Study of the
(36) Chidambaram, A.; Stylianou, K. C. Electronic metal-organic reaction mechanism of the transesterification of triglycerides catalyzed
framework sensors. Inorg. Chem. Front. 2018, 5 (5), 979−998. by zinc carboxylates. J. Mol. Catal. A: Chem. 2013, 377, 29−41.
(37) Gumilar, G.; Kaneti, Y. V.; Henzie, J.; Chatterjee, S.; Na, J.; (55) Chen, L.; Zhan, W.; Fang, H.; Cao, Z.; Yuan, C.; Xie, Z.;
Yuliarto, B.; Nugraha, N.; Patah, A.; Bhaumik, A.; Yamauchi, Y. Kuang, Q.; Zheng, L. Selective Catalytic Performances of Noble Metal
General synthesis of hierarchical sheet/plate-like M-BDC (M = Cu, Nanoparticle@MOF Composites: The Concomitant Effect of
Mn, Ni, and Zr) metal-organic frameworks for electrochemical non- Aperture Size and Structural Flexibility of MOF Matrices. Chem. -
enzymatic glucose sensing. Chemical Science 2020, 11 (14), 3644− Eur. J. 2017, 23 (47), 11397−11403.
3655.
(38) Wang, X.; Wu, X.-L.; Guo, Y.-G.; Zhong, Y.; Cao, X.; Ma, Y.;
Yao, J. Synthesis and Lithium Storage Properties of Co 3 O 4
Nanosheet-Assembled Multishelled Hollow Spheres. Adv. Funct.
Mater. 2010, 20 (10), 1680−1686.
(39) Liu, N. L.; Dutta, S.; Salunkhe, R. R.; Ahamad, T.; Alshehri, S.
M.; Yamauchi, Y.; Hou, C. H.; Wu, K. C. ZIF-8 Derived, Nitrogen-
Doped Porous Electrodes of Carbon Polyhedron Particles for High-
Performance Electrosorption of Salt Ions. Sci. Rep. 2016, 6, 28847.
(40) Jiao, L.; Wang, Y.; Jiang, H. L.; Xu, Q. Metal-Organic
Frameworks as Platforms for Catalytic Applications. Adv. Mater. 2018,
30 (37), No. 1703663.
(41) Liao, Y.-T.; Matsagar, B. M.; Wu, K. C. W. Metal-Organic
Framework (MOF)-Derived Effective Solid Catalysts for Valorization
of Lignocellulosic Biomass. ACS Sustainable Chem. Eng. 2018, 6 (11),
13628−13643.

6550 https://doi.org/10.1021/acssuschemeng.0c08012
ACS Sustainable Chem. Eng. 2021, 9, 6541−6550

You might also like