0% found this document useful (0 votes)
93 views30 pages

Upcycling and Catalytic Degradation of Plastic Wastes

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
93 views30 pages

Upcycling and Catalytic Degradation of Plastic Wastes

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 30

ll

OPEN ACCESS

Review
Upcycling and catalytic
degradation of plastic wastes
Qidong Hou,1 Meinan Zhen,1 Hengli Qian,1 Yifan Nie,1 Xinyu Bai,1 Tianliang Xia,1
Mian Laiq Ur Rehman,1 Qiushi Li,1 and Meiting Ju1,*

SUMMARY
Various recycling technologies have been developed to deal with plas-
tic problems, but they face considerable economic and technological
challenges in practice. An attractive alternative is upcycling, which
aims to dig out the embedded value to incentivize large-scale valoriza-
tion of plastic wastes. The degradation of nonrecoverable plastic
wastes is another necessity to treat the omnipresent pollution. This re-
view presents an overview on the conversion of plastic wastes toward
value-added products and the catalytic degradation of nonrecover-
able plastic wastes. Based on an examination of traditional recycling
technologies and products, we summarize the state-of-the-art design
and development of plastic conversion to high-value and high-perfor-
mance fuels, chemicals, and materials. Subsequently, we highlight the
advances in catalytic degradation of plastics to environmentally
benign or degradable products and mineralization into carbon dioxide
and water. We conclude with our perspective on the ongoing chal-
lenge and opportunities.

INTRODUCTION
Since the appearance of the first synthetic polymer in 1907, various plastics have
been increasingly manufactured and used in ever-expanding fields because of their
low cost, durability, safety, and processability.1 Nowadays, plastics-based materials
have become ubiquitous and indispensable in modern society, with a wide scope
and range of applications that cannot currently be displaced by other materials.2
Along with an excessive dependence on plastic products, the end-of-life treatment
of post-consumer plastic wastes has become a huge problem because plastics are
designed to be robust and durable. As of 2015, 6,300 Mt of plastic wastes had
been cumulatively generated; of which, 79% entered landfills or accumulated in
the natural environment, and 12% was incinerated.3 Only 9% of plastic wastes
have been recycled, and most of the recycled plastic is down-cycled toward less-
recyclable and low-value products, such as plant pots and garden furniture.4,5

Because of this externality, global plastic pollution, as a planetary boundary threat,


had already formed before people began to realize its enormity, complexity, and ur-
gency.6,7 Plastic wastes are omnipresently and increasingly observed in oceans,
lakes, rivers, soils, sediments, the atmosphere, and animal biomass.8–12 Microplastic
(%5 mm) and nanoplastic (%100 nm) pollution originated from both direct emission
of ‘‘microbeads’’ and ‘‘micro-exfoliates’’ (primary microplastics) contained in house-
1National & Local Joint Engineering Research
hold cosmetic products into household wastewater and from the breakdown of
Center of Biomass Resource Utilization, College
larger plastic waste into small plastic pieces (secondary microplastics) via photoox- of Environmental Science and Engineering,
idation under solar irradiation, physical crushing, and biodegradation in the natural Nankai University, Tianjin 300350, China
environment.13–18 Microplastics and nanoplastics can be ingested by various organ- *Correspondence: [email protected]
isms, and nanoplastics can even accumulate in plants,16,19 ultimately resulting in https://doi.org/10.1016/j.xcrp.2021.100514

Cell Reports Physical Science 2, 100514, August 18, 2021 ª 2021 The Author(s). 1
This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
ll
OPEN ACCESS
Review

their enrichment via food webs. Plastic debris can function as a medium to compile
and spread hydrophobic organic pollutants, heavy metals,20 and diseases.20
Although the direct toxicological effect of plastics on human health has not been
substantiated, the ever-increasing emission of plastics is destined to generate mul-
tiple harmful effects.5,21 For example, microplastics have increasingly entered into
the human food system, such as seafood,22 tea,23 and vegetables,21 as a nonnegli-
gible threat to food safety and agricultural sustainability.24 Moreover, microplastics
have been detected in human placentas.25 In addition, global green house gas
(GHG) emissions from the plastic life-cycle is expected to rise from 1.7 Gt of carbon
dioxide (CO2) equivalent in 2015 to 6.5 Gt in 2050 under current trends, contributing
significantly to climate change.26

The improper management of plastic wastes not only contaminates the environment
but also signifies an immense loss of economic value.27 Globally, the energy savings
for plastic waste recycling are estimated to be 3.5 billion barrels of oil, equivalent to
about $176 billion dollars.28 Although several recycling technologies have been
investigated, they suffer universally from low benefits, high costs, and secondary
pollution, leading to limited practical applications.29 Therefore, the development
of cost-effective, environmentally friendly, and efficient approaches to valorize plas-
tic wastes into value-added products is indispensable to prevent their dissemination
into natural environment. In addition, the development of effective catalytic-degra-
dation technologies is essential for treating nonrecoverable plastic wastes.

We begin this review by providing a brief assessment on possible options to address


the plastic problems. In combination with the development of recycling technolo-
gies, we summarize recent advances in upcycling to valuable chemicals, fuels, and
materials. Then, we focus on the catalytic degradation of unrecoverable plastic
wastes. We do not provide a comprehensive discussion on other related important
aspects, such as traditional recycling technologies,28,30–32 biotechnological degra-
dation,29,33 production of polymers from renewable resources,34 and the develop-
ment of recyclable and degradable polymers,35,36 because excellent reviews could
be found elsewhere.

OPTIONS TO ADDRESS THE PLASTICS PROBLEM


Most plastic wastes fall into four top categories (Figure 1A): polyester, polyolefin,
polyvinyl chloride (PVC), and polystyrene (PS).37 Polyethylene terephthalate (PET)
is the most widely used polyester, with a global annual production of 70 million
tons (Mt) for use in textiles, packaging, carpeting, and single-use beverage bot-
tles.38,39 Polyolefins, including polyethylene (PE) and polypropylene (PP), with
annual production of approximately 138 and 80 Mt, respectively, account for 57%
of the plastic content of municipal solid waste.3,40 Among PEs, high-density polyeth-
ylene (HDPE) is used extensively for electrical insulators, toys, bottles, pipes, and
films, whereas low-density polyethylene (LDPE) is mainly used in devices, food pack-
aging, and plastic wrap.28 As the cheapest plastic, PVC is one of the most problem-
atic wastes for the environment because of the emission of phthalate plasticizers and
chlorine-containing hydrocarbons (dioxins) during its end-of-life treatment.37
Detailed information on the production, properties, and application of other plastics
can be found in previous reviews.28,41 Among the various plastic wastes, only PET
and HDPE are routinely recycled, but the reclamation rate is far from satisfactory.39

The 6R principle (reduction, repair, reuse, recover, remanufacturing, and recycling,42


with another version as rethink, refuse, reduce, reuse, recycle, and replace33) of the

2 Cell Reports Physical Science 2, 100514, August 18, 2021


ll
OPEN ACCESS
Review

C D

Figure 1. Options to address the plastic problems


(A) Main types of plastic waste.
(B) Process of plastic-waste generation.
(C) Traditional recycling technologies.
(D) Upcycling toward value-added products.
(E) Degradation of plastic wastes.

circular economy offer a basic framework to seek solutions for the plastics problem.
In principle, cutting down the use of plastic materials (reduction) and substituting
alternative materials for the plastic (Figure 1B) would prevent the generation of plas-
tic wastes and, in addition, relieve the excessive dependence on depletable re-
sources. Despite considerable endeavors, currently, sophisticated plastic materials
will continue to dominate for the long term, and many of them cannot be replaced
with other materials because of the cost and the disparity in performance.43,44 For
degradable and recyclable materials, it is difficult to concurrently satisfy the tough

Cell Reports Physical Science 2, 100514, August 18, 2021 3


ll
OPEN ACCESS
Review

requirements for material properties, cost, and degradability or recyclability. The


design and development of degradable polymers, which undergo triggered degra-
dation under environmentally authentic conditions once they are disposed of in the
natural environment, must still guarantee long-term durability during shelf and ser-
vice life, which is extremely challenging.35 For instance, polylactic acid (PLA, also
termed polylactide), one of most widely used degradable materials at present, is
basically non-degradable in seawater.45

Numerous recycling processes (Figure 1C), including primary (closed-loop) recy-


cling, secondary (mechanical) recycling, tertiary (chemical) recycling, and incinera-
tion have been established to treat post-consumer plastic wastes. Both primary
and secondary recycling are based on mechanical processes, including sorting,
grinding, washing, remodeling, and extruding.28 Primary recycling refers to pro-
cessing of discarded plastic to make fresh products that are returned to the
same or similar-quality applications. That process can only treat near-pristine plas-
tic of a known origin.28,30 Secondary recycling refers to reprocessing of plastic
wastes into products for uses that are different from the initial material through
mechanical processes.28 Because processing under harsh conditions (high temper-
ature and mechanical force) unavoidably results in the degradation of polymers,
the recovered low-grade plastics with inferior optical, thermal, and mechanical
properties can only be employed in applications with less-engineering require-
ments than that of the original application46, such that the mechanical recycling
is also termed ‘‘downgrading’’ or ‘‘downcycling.’’28 Currently, mechanical recycling
is only used commercially for the recovery of PET and PE.27 The ultimate fate
of mechanically recovered plastic products is still landfilling, disposal, or incinera-
tion because of performance deterioration of the regenerated materials.1 In addi-
tion, the mechanical recycling processes usually emit several volatile organic
compounds.47

Tertiary recycling refers to recycling of plastics through chemical routes, such as py-
rolysis, gasification, and depolymerization.28 Pyrolysis (also called thermolysis, ther-
mal cracking, catalytic cracking, and liquefaction) converts plastics to gases, liquids,
and waxes under high temperatures, either in the absence of a catalyst or in the pres-
ence of a catalyst (catalytic pyrolysis).31 Gasification aims to produce gas products
under high temperatures with a gasification agent, such as air or oxygen.48 Because
the polymeric molecules of the plastic typically have a high molecular weight
(5,000–5,000,000 g mol–1) with robust chemical bonds, their degradation into a
myriad of small molecules via multiple side reactions seems to be inevitable in tradi-
tional pyrolysis processes, resulting in poor control of the reaction pathway and low
selectivity for desirable products.49 Avoiding side reactions and narrowing product
distributions are fundamentally important for traditional pyrolysis processes. Depo-
lymerization refers to converting the polymers to well-defined monomers or oligo-
mers via the disruption of certain chemical bonds.31 For a detailed discussion of
these process descriptions, the reader is referred to a more specialized review.31
Chemical recycling is ill-suited for many kinds of plastic wastes that are composed
of robust chemical linkage, such as polyolefins.43 Even for polymers with tractable
chemical bonds, the recovery of monomers from plastic wastes is often costly, en-
ergy intensive, and incompatible with complex mixtures.50 Because real-world plas-
tic wastes generally contain plasticizers, inks, dyes, and adhesives39 and are mixed
with other polymers or contaminated by other wastes, chemical recycling requires
time- and energy-consuming separation and purification steps, thus leading to a
further decrease in overall economic feasibility or resulting in reduced performance
of the regenerated products.

4 Cell Reports Physical Science 2, 100514, August 18, 2021


ll
OPEN ACCESS
Review

In comparison with recycling technologies, incineration is a more all-purpose tech-


nology, which has been routinely employed for end-life treatment of mixed or
contaminated wastes at sizable scales. Although incineration can recover partial en-
ergy in the form of heat from mixed wastes without the need for separation, it is an
unfavorable process in view of the overall energy loss relative to recycling,51 high
GHG emissions (5 to 10 tons of CO2 per ton of plastic during the entire life cycle),31
and toxic combustion products.26,52

Frustratingly, the above strategies cannot end the environmental crisis associated
with anthropogenic plastics. A state-of-the-art scenario analysis shows that imple-
menting the most aggressive solutions from 2016 to 2040 through the use of
currently available knowledge and technologies, including collection, disposition,
recycling, reduction, and substitution could reduce plastic pollution by 40% relative
to 2016, but 710 Mts of plastic waste will still be discharged inescapably into aquatic
and terrestrial ecosystems, accompanied by the emission of large quantities of
GHGs from plastic manufacturing and unsound waste-management activities.8
Therefore, extraordinary efforts are indispensable to achieve a substantial reduction
in global plastic emissions.53

To overcome the inherent defect of conventional strategies, upcycling, which em-


phasizes the exploration of an inherent value in plastic wastes (Figure 1D), has
been developed as a complementary and more attractive option.39,54 Until recently,
there has not been an accurate definition of the concept of ‘‘upcycling.’’ The terms
‘‘recycling’’ and ‘‘upcycling’’ have been used indiscriminately in some publications
because some routes do have some overlaps and many upcycling processes are
developed based on recycling processes. Comparatively, the former emphasizes
‘‘closed-loop’’ for the plastic materials themselves, whereas the latter will be an
open-loop processes toward multiple profit streams. In this review, we define upcy-
cling as the conversion of low-cost and abundant plastic wastes into high-perfor-
mance fuels, chemicals, and materials via digging out the embedded value present
in the form of carbon, hydrogen, chemical, energy, or macromolecular structure.
Ideal, sustainable, or circular plastic-waste refineries55 may be established to afford
fuels, chemicals, and materials, similar to the well-applied oil refinery and preliminar-
ily established biorefinery roadmap.56 Moreover, upcycling processes may provide
new approaches to handle real-world plastic wastes, in particular. those cannot be
subjected to thermomechanical recycling. Therefore, there is no doubt that upcy-
cling of plastic waste would simultaneously contribute to the mitigation of solid
waste contamination and to the production of high-value products, hence, signifying
considerable economic and scientific opportunities.57

Both recycling and upcycling are designed for the valorization of post-consumer
plastic wastes to prevent the emission of plastic wastes into the natural environment,
but they cannot deal with nonrecoverable plastic wastes. In fact, there are vast and
diverse plastic wastes that cannot be collected and used under current economic
and technical condition, such as plastic fragments mixed with sludge and plastic
debris dispersed in the natural environment. Commercially, running-water treat-
ment technologies can remove microplastics from wastewater via skimming, sedi-
mentation, and tertiary filtration, but most plastics without substantial degradation
are still retained in sewage sludge.58–60 Environmental degradation of plastic
wastes, including abiotic degradation and biodegradation behaviors, depends on
many complex factors, including physical properties and molecular structures,61,62
and generally suffers from low efficiency.63 Because of the strong chemical inertness
of plastic wastes, natural degradation can take decades and continue to emit

Cell Reports Physical Science 2, 100514, August 18, 2021 5


ll
OPEN ACCESS
Review

hazardous organic pollutants during that time.64,65 Successful enzymatic or microbi-


al degradation has largely concentrated upon a few types of plastic that contain hy-
drolyzable bonds.33 Although considerable achievement has been reported on
biotechnological degradation, its effectiveness under natural conditions remains
to be validated.33 Therefore, the development of effective catalytic degradation
technologies, which could completely mineralize nonrecoverable plastic waste
into CO2 and water or degrade them to environmentally benign or degradable inter-
mediates (Figure 1E), is of high importance for the remediation of plastic pollution.

UPCYCLING TO CHEMICALS
Catalytic depolymerization to monomers
Catalytic depolymerization to monomers, which is also termed chemical recycling to
monomer,31 is an important form of chemical recycling.30 It is different from mechan-
ical recycling because the catalytic depolymerization of initial monomers enables
the production of recovered plastic with properties that are the same as virgin plas-
tic, if purified monomers can be obtained.39 Under the framework of upcycling, both
monomers and partially deconstructed polymers are important feedstocks for the
development of new processes toward value-added chemicals, fuels, and materials.
Catalytic depolymerization to monomers relies heavily on the presence of fragile
chemical bonds in the skeleton of the polymers. At present, catalytic depolymeriza-
tion to monomers predominantly concentrates upon polyesters, in particular PET,
because the chemolysis of esters is relatively facile.

Various catalytic depolymerization methods have been investigated for the conver-
sion of PET to monomers. The primary processes include hydrolysis with water to
terephthalic acid (TPA) and to ethylene glycol (EG) under neutral, acidic, or basic con-
ditions (Figure 2A); alcoholysis with alcohol (methanol, ethanol, etc.) to dialkyl tere-
phthalate and EG; glycolysis with excess glycols (such as EG, diethylene glycol
[DEG], propylene glycol [PG], polyethylene glycol, 1,4-butanediol. and hexylene gly-
col) toward bis(hydroxylethylene) terephthalate (BHET or other corresponding es-
ters) via a transesterification reaction; and aminolysis with amines (or ammonolysis
with ammonia) toward corresponding diamides of TPA and EG;66,67 among which,
some depolymerization processes are also termed solvolysis.31 The reader can find
systematic summaries of these processes in specialized reviews mentioned above.
These methods can also be used for the depolymerization of other polyester plastics,
such as polybutylene terephthalate (PBT).68 Glycolysis is also the predominant cata-
lytic depolymerization method for polyurethanes (PUR).69 The harsh reaction condi-
tions, the use of excess solvents and depolymerizing agents, as well as the formation
of by-products are the main obstacles to commercial deployment of these processes.

Catalytic depolymerization back to the original monomers is rather difficult for many
extant plastics, which are composed of robust chemical linkages because of the
absence of fragile chemical bonds. For example, prohibitive amounts of energy
are required to depolymerize polyolefins, such as PE and PP, to their original mono-
mer subunits, along with low selectivity.43 Similarly, pyrolysis of PS usually generates
complex products with low selectivity.75 Among polyamides (PAs), only nylon-6,
which is synthesized from the seven-membered ring ε-caprolactam, is suitable for
depolymerization back to monomers with relatively high yield and selectivity.30
The depolymerization of most polyethers is difficult because of the challenge of acti-
vating unstrained ethers, whereas poly(tetrahydrofuran), an amorphous soft material
widely used as a synthetic component in various industrial and consumer applica-
tions, is relatively amenable to catalytic depolymerization.76

6 Cell Reports Physical Science 2, 100514, August 18, 2021


ll
OPEN ACCESS
Review

Figure 2. Catalytic depolymerization of plastics to monomers


(A) Depolymerization of PET via hydrolysis or trans-esterification. 66
(B and C) Hydrogenolysis of PET over (B) C/MoO 2 46 and (C) Ru catalysts. 70–72
(D) Hydrogenolysis of aromatic plastics over a Ru/Nb 2 O 5 catalyst. 73
(E) Hydrogenolysis of PET using a [Ru]/KOt Bu catalyst with dimethyl sulfoxide (DMSO) as the solvent. 74

Similarly, biotechnology offers promising paths for the depolymerization of poly-


mers to well-defined oligomers or monomers for recycling and upcycling but is
also limited to the few plastics with hydrolyzable esters or amides bonds, such as
PET, PA, and selected types of PUR.33 Recently, enzymatic hydrolysis achieved
remarkable progress: for instance, the engineered PET depolymerase, which is ob-
tained by computer-aided enzyme engineering, has afforded 100% PET conversion
within 10 h, with a PTA productivity as high as 16.7 g L 1 h 138. Another study shows
that nanoscopically dispersing enzymes into polyesters during their production
could greatly improve substrate accessibility, thus, allowing near-complete depoly-
merization of polyesters.77

Catalytic hydrogenolysis to chemicals


Hydrogenolysis, a specific form of depolymerization, refers to disrupting chemical
bonds, in particular, C–C bonds, with the assistance of hydrogen (H2).31 In some

Cell Reports Physical Science 2, 100514, August 18, 2021 7


ll
OPEN ACCESS
Review

cases, plastics can be selectively deconstructed through hydrogenolysis to short-


chain products with values that are substantially higher than the fully deconstructed
monomers.44 At present, upcycling of plastic wastes to value-added chemicals via
direct hydrogenolysis mainly concentrates on PET and PE. Compared with the depo-
lymerization processes based on hydrolysis or transesterification, catalytic hydroge-
nolysis has offered more accessible and promising options for the conversion of PET
wastes to valuable chemicals and the drop-in integration of plastic valorization with
well-established industrial processes toward an ideal circular economy. The use of
the carbon-supported single-site molybdenum-dioxo (C/MoO2) catalyst enables se-
lective depolymerization of PET to TPA and ethylene (Figure 2B) under solvent-less
conditions at 260 C in the presence of 1 atmosphere of H2,46 which is rarely achieved
in traditional chemical and biological depolymerization processes. The reaction over
C/MoO2 catalyst proceeds via retro-hydroalkoxylation/b-C–O scission, followed by
hydrogenolysis of vinyl benzoate. Moreover, high TPA yields are also obtained from
waste beverage-bottle PET and PET/PP mixtures (simulation of bottle and cap,
respectively), and the catalyst is recyclable without losing activity.

Selective hydrogenolysis of PET to 1,4-benzene dimethanol and EG (Figure 2C) un-


der mild temperatures has been achieved over a series of ruthenium (Ru)-based
molecular catalysts,71,72 using anisole, tetrahydrofuran (THF), or their mixture as sol-
vent. These molecular catalysts are also capable of selectively depolymerizing PLA,
polycaprolactone (PCL), poly(bisphenol A carbonate) (also termed bisphenol A pol-
ycarbonate [BPA-PC]), and PBT to the corresponding diols.72 The large-scale appli-
cation of expensive, air-sensitive, potentially toxic Ru catalysts with complex ligands
is unrealistic because of the difficulty of catalyst recycling, despite their high activity
and selectivity.46

Converting PET back to p-xylene and methylbenzene (Figure 2D) has been achieved
via catalytic hydrogenolysis over a Ru/Nb2O5 catalyst in water.73 This catalytic sys-
tem also enables the hydrogenolysis of other aromatic plastic wastes with interunit
C–O and/or C–C linkages toward arenes (75%–85% yield), including PS, poly(p-phe-
nylene oxide) (PPO), BPA-PC, and mixed aromatic plastic wastes. In contrast, Pd-
and-Pt-based catalysts convert PET to ring-saturated products, whereas Ru/ZrO2,
Ru/TiO2, and Ru/HZSM-5 catalysts convert PET to mixtures ranging from arenes to
ring-saturated products with low selectivity. Compared with other catalysts, the
Ru/Nb2O5 catalyst is selective in tailoring the C–O bond and the sp2–sp3 bond,
that is, Caromatic–C connected with aromatic rings, without hydrogenating aromatic
ring, because of the presence of sub-nano Ru clusters and Lewis/Brønsted acid sites
on Nb2O5. When treating mixed plastics, the Ru/Nb2O5 catalyst loses some activity
(28%) in the second run because of carbon deposition, but it is restored completely
after calcination in air, followed by reduction with H2.

Catalytic hydrogenation of robust PAs is difficult because they are resistant to most sol-
vents because of the multiple, strong intermolecular hydrogen bonding interactions
within the polymer chains. Hydrogenative depolymerization of nylon-6 to amino alco-
hols (Figure 2E) has been achieved by a combination of a ruthenium pincer ([Ru]/KOtBu)
catalyst in DMSO.74 The use of DMSO as a solvent is crucial to disrupt hydrogen
bonding and, in addition, to maintain the uncoordinated state of the metal center,
whereas metal-ligand cooperativity is important for the hydrogenation of amide bonds
through an outer-sphere mechanism. This catalytic system can also catalyze the hydro-
genation of PUR to diol, diamine, and methanol. The diol and diamine can be used to
produce original PA through a dehydrogenative coupling process using the same Ru
catalyst, providing a potential route for the recycling of nylons.

8 Cell Reports Physical Science 2, 100514, August 18, 2021


ll
OPEN ACCESS
Review

PET, except for H2, silanes have also been investigated for reductive depolymeriza-
tion of plastic wastes. For example, PLA, PCL, and PBT can be successfully converted
into value-added compounds, including p-xylene, 1,2-propanediol, 1,6-hexanediol,
and THF, with a Zn(OAc)2 $ 2H2O catalyst using (EtO)2MeSiH or PhSiH3 as the reduc-
tant.78 The advantages of this catalystic system are its low cost and the good reus-
ability of the catalyst, but silanes are very expensive.

Other routes to useful chemicals


Direct hydrogenolysis of polyolefins, including PE and PP, generally produce mixed
alkanes with a wide molecular distribution, instead of well-defined monomers, even
when elaborately designed catalytic systems are used, as will be discussed in the sec-
tion of upcycling to fuels. The consumption of expensive H2, which essentially origi-
nates from non-renewable fossil fuel resources, is a primary barrier to the application
of hydrogenolysis technologies. Tandem catalysis, which refers to integrating multi-
ple reaction steps into one-pot catalytic systems in an appropriate sequence through
precise regulation of active sites, the chemical environment, and the reaction condi-
tions, offers a promising strategy to inhibit unwanted side reactions to tailor a reac-
tion pathway and then achieve selective, efficient conversion of plastic waste to
target products. Recently, upcycling of PE to long-chain alkylaromatics (Figure 3A)
has been achieved by tandem hydrogenolysis/aromatization over a commonly
used heterogeneous catalyst, platinum nanoparticles supported on g-alumina (Pt/
Al2O3), without consuming the external hydrogen.43 In contrast to pyrolysis, which
generates complex, low-value mixtures under harsh condition (>400 C), the tandem
catalytic system produces high yields (up to 80 wt%) of low-molecular-weight liquid/
wax products with high selectivity for valuable long-chain alkylaromatics and alkyl-
naphthenes (average carbon atom number of 30, dispersity [Ð] = 1.1) from various
PE grades under solvent-less conditions at a moderate temperature (280 C). The
conversion of two commercial-grade PE samples with higher-molecular-weight,
LDPE plastic bag and HDPE water-bottle caps produces overall liquid yields of 69
wt% and 55 wt% with alkylaromatic selectivities of 44 mol% and 50 mol%, respec-
tively, validating its applicability to real-world plastic wastes. During the reaction, hy-
drogenolysis of PE chains can consume H2 and then reduce the thermodynamic
values (DG and DH) of aromatization, whereas aromatization generates H2 to induce
the hydrogenolysis (Figure 3B). The liquid alkylaromatics can be used as feedstocks
to produce various products of daily life, such as surfactants, lubricants, refrigeration
fluids, and insulating oils. In addition, theoretical simulation suggests that the tan-
dem olefin metathesis and double-bond isomerization may selectively convert PE
waste with excess ethylene into propylene products that are worth about
$0.200.30/mol higher than an ethylene feedstock.79

Taking advantage of existing refinery technology and infrastructure is crucial to


exploring effective recycling and upcycling routes. For instance, the refinery-waste
catalyst, which is discarded as waste during the industrial-fluid catalytic-cracking
process is employed to convert PP, and the same aromatics content is obtained
with fresh catalyst with considerably inhibited coke deposition.81 The improved
aromatization and pre-cracking activity is attributed to the deposition of Fe, Ni,
and V onto the catalyst during the fluid catalytic cracking. Meanwhile, the lifetime
of the catalyst is improved because of the decreased coke deposition, which is
attributed to the block of micropore structure in the zeolite material as well as to
the absence of acid sites.

A tandem-catalysis strategy has also been successfully used for the conversion of
PET to benzene, p-xylene, and methylbenzene (BTX) without using external

Cell Reports Physical Science 2, 100514, August 18, 2021 9


ll
OPEN ACCESS
Review

Figure 3. Upcycling of PE and PET through a strategy of tandem catalysis


(A and B) Schematic illustration of PE upcycling to long-chain alkylaromatics via tandem
hydrogenolysis/aromatization (A) and a reaction pathway and mechanism for the conversion of PE
to alkylaromatics and alkylnaphthenes (B). 43,57 (A) was adapted with permissions, 57 copyright 2020,
AAAS.
(C) H 2 -free conversion of PET to arenes. 80

hydrogen sources (Figure 3C), as an alternative approach to hydrogenolysis.80


Compared with the hydrogenolysis of PET over a Ru/Nb2O5 catalyst at 200 C in
the presence of 0.3 MPa H2, H2-free conversion of PET back to BTX in the same cat-
alytic system was achieved by unearthing hidden structural hydrogen under rela-
tively harsh reaction conditions (220 C), affording total monomer yields up to
91.3% with quantitative selectivity to BTX (>99.9%). The Ru/Nb2O5 catalyst could
catalyze three tandem steps, including PET hydrolysis, EG reforming to H2, and hy-
drogenolysis of C–O/C–C bonds, along with partial decarboxylation reaction.

The application of other catalytic technologies in plastics upcycling may also pro-
duce useful products. For example, PE waste can be used to synthesize plasticizers
for PLA.82 Under microwave radiation, both HDPE and LDPE are selectively oxidized
by HNO3 to dicarboxylic acid mixtures, including succinic, glutaric, and adipic
acid.82,83 The reaction of dicarboxylic acid mixtures with 1,4-butanediol and crotonic
acid delivers plasticizers. The resulting plasticizers can be covalently attached to PLA
by a reactive extrusion process, improving the strain at break to 142% and
decreasing the glass-transition temperature by 10 C. The integrated processing of

10 Cell Reports Physical Science 2, 100514, August 18, 2021


ll
OPEN ACCESS
Review

polyoxymethylene polymer waste with biomass-derived diols, such as 1,3-propane-


diol over an acid catalyst, affords cyclic acetals as promising solvents, fuel additives,
pharmaceutical intermediates, and polymeric monomers.84

UPCYCLING TO FUELS
The elementary compositions of many plastics, in particular the C and H content in
polyolefins are close to petroleum-derived hydrocarbons. As a consequence, the
calorific values of plastic wastes are comparable to currently used liquid fuels.85 In
addition, plastic wastes are also considered as alternative hydrogen-rich energy
feedstock because of the relatively high hydrogen content (approximately 8–14 wt
%).49 Although pyrolysis and gasification have been extensively investigated to
convert plastic wastes to fuels, they yield only complicated mixtures, which require
tedious separation, purification, and upgrading process to provide commercial fuel
products. In contrast, upcycling to high-performance fuels, including H2 and liquid
alkanes with a narrow molecular-weight distribution has shown great promise for
the valorization of plastic wastes.

Hydrogen
Photoreforming is a solar-driven technology that uses the energy of sunlight to
convert organics and water to value-added products, such as H2 and small organic
molecules.86 Most research on photocatalytic H2 production is based on the decom-
position of an unrealistic sacrificial agent, such as triethanolamine into H2, instead of
overall water splitting, whereas earlier photoreforming of organic waste universally
focuses on simple ‘‘model waste’’ substrates, such as ethanol, glycerol, or soluble
sugars. For many common organic wastes, the overall photoreforming process is
nearly energetically neutral and more favorable than water splitting.86 Therefore,
photoreforming has provided a promising approach to simultaneously deal with
organic solid wastes and energy challenges with the reduction of anthropogenic
GHG emissions.87 Generally, the photoreforming of hydrocarbon-chain-based plas-
tics (non-polar polymers), is challenging because of the difficulty of breaking down
highly stable C–C bonds,86 despite the relatively high hydrogen content. In contrast,
the plastics that contain oxygenated, polar groups or esters (polar polymers), such as
PET, PUR, and PLA, are more amendable feedstocks.

The photoreforming of actual plastic was first attempted in 1981 using platinized
TiO2 as a photocatalyst in an NaOH aqueous solution (5 M), with low H2 evolution
rates.88 The visible-light-driven, noble, metal-free photoreforming of plastic has
been achieved by combining CdS/CdOx quantum dot catalyst with 10 M NaOH so-
lution (Figure 4). This catalytic system attains maximal H2 production rates of 64.3 G
14.7, 3.42 G 0.87, and 0.85 G 0.28 mmol H2 gCdS 1 h 1 from PLA, PET, and PUR,
respectively.89 When using realistic PET water bottle as the feedstock, it still affords
H2 production rates of 4.13 G 0.40 mmol H2 gCdS 1 h 1 with a rate of conversion of
5.15% G 0.72% and an external quantum yield of 2.17% G 0.38%. The superior cat-
alytic performance of CdS/CdOx is attributed to the excellent visible light absorp-
tion ability, high dispersity, and high activity under strong alkaline condition as
well as the partial dissolution and hydrolysis of plastics to monomers induced by
NaOH. The strong alkaline condition leads to the in situ formation of thin oxide/hy-
droxide layer on the surface of CdS, which could not only prevent the photocorro-
sion of CdS but also improve the photocatalytic activity.90 However, the H2 cannot
be completely released, and the overall conversion does not outperform 40%,
with the formation of organic molecules, such as formate, acetate, and pyruvate,
which undergo slow photoreforming.

Cell Reports Physical Science 2, 100514, August 18, 2021 11


ll
OPEN ACCESS
Review

Figure 4. Photoreforming of plastics to H2


(A) Schematic illustration of the photoreforming process of plastics over a CdS/CdOx quantum-dot
photocatalyst in an alkaline aqueous solution. 89 Adapted with permissions, 89 copyright 2018, Royal
Society of Chemistry.
(B–D) Conversion of PLA (B), PET (C), and PUR (D) to organic products. 89

To displace the noble metal and Cd-based catalysts, cheap and nontoxic carbon
nitride has been investigated for photoreforming.91 After pretreatment with an
NaOH solution (stirring at 40 C in the dark for 24 h), 72% of PLA is converted to
lactate, whereas 62% of EG is released from PET. The cyanamide-functionalized car-
bon nitride (CNx) coupled with a nickel phosphide (Ni2P) provides two and four times
less H2 yield than those with CdS/CdOx for PLA and PET, respectively, with oxidation
products similar to those over CdS/CdOx. The additional electron-transfer process
from the light absorber CNx to the cocatalyst Ni2P is one of the main limiting factor
for the activity. The catalytic system can convert microplastics (polyester microfibers)
and food-contaminated plastic without compromising catalytic activity when the
reactor is upscaled from 2 to 120 mL.

These photoreforming studies represent useful proofs-of-concept, but their applica-


tion faces a series of economic and technologic hurdles. It is estimated that the
economically and environmentally feasible running of photoreforming requires a
50 times greater H2 evolution rate (>0.004 mol H2 gsubstrate 1 h 1, conversion of
>50% per day, and external quantum yield of >50%) and greater durability of the
photocatalyst (lifetime > 1 year).86 Currently, the knowledge of the mechanism of
oxidation half-reaction, the rate-limiting step for photoreforming, is still limited,
and the selective oxidation of plastic to value-added chemicals has not been proven.
Strong alkaline conditions are crucial for partial solubilization and hydrolysis of solid
feedstock both in pretreatment and photoreforming processes and can greatly
improve photoreforming efficiency. However, large-scale use of alkaline solution
is not viable in terms of both cost and energy returned on energy invested, even
assuming it can be recycled 60 times. The design of more-efficient and stable

12 Cell Reports Physical Science 2, 100514, August 18, 2021


ll
OPEN ACCESS
Review

photocatalysts and co-catalysts, the improvement of substrate-catalyst contact, the


screening of suitable solvent,92 the development of pretreatment methods, and an
examination of useful oxidation products are needed to promote the establishment
of economically, environmentally, or energetically favorable photoreforming
processes.

Selective, rapid, and one-step upcycling of real-world plastic wastes, including milk
containers (HDPE), plastic bags (LDPE), food wraps (PP), and plastic foam (PS) to high
purity H2 and high-value carbon materials, predominantly in the form of multiwalled
carbon nanotubes (MWCNTs) has been achieved by microwave-initiated catalytic
deconstruction.49 Composite iron oxide and aluminum oxide (FeAlOx) functions as
both the microwave susceptor and the catalyst to initiate a rapid conversion of plas-
tics under solvent-less conditions, obtaining excellent H2 yields of 55.6, 51.4, and
26.9 mmol gplastic 1 from mechanical, pulverized HDPE, PP, and PS (theoretical
yield, 71.4, 71.4, and 38.4 mmol gplastic 1), with negligible liquid and carbon yield
up to 0.70, 0.60, and 0.80 g $ gplastic 1, respectively. Combined with other gas prod-
ucts, near-quantitative hydrogen (97%) has been extracted within 90 s. In contrast,
both the use of carbon materials as a catalyst under microwave conditions and the
use of FeAlOx as the catalyst in traditional pyrolysis processes produce low H2 yields.
The effective conversion is due to the following reaction mechanism: (1) the selective
generation of heat over the catalyst; (2) heat is transferred from the catalyst to the
plastic; (3) the microwave initiates the scission of the C–H bonds over the catalyst,
forming H2 and carbon; (4) and the partial-carbon species reacts with the FeAlOx
to form Fe3C, which assists in the generation of the MWCNTs. Unlike conventional
heating processes, the preferential heating of the catalyst could promote the
desorption of hydrogen from active sites and avoid fundamental side reactions,
thus leading to the selective conversion of plastics toward target products.

Alkanes
Catalytic hydrogenolysis can convert PE and PP into alkane products, including lubri-
cants, waxes, diesels, and light-hydrocarbon gases, but these processes generally
suffer from a lack of product control and consumption of expensive H2.93,94 Recently,
upcycling of HDPE toward diesel and lubricant-range alkanes with high yield (79%)
and narrow distribution (from C9 to C18) has been achieved over an ordered, meso-
porous shell/active site/core catalyst (mSiO2/Pt/SiO2), a chemically and thermally
robust inorganic material.44 In contrast, the Pt nanoparticles supported on silica
spheres (Pt/SiO2) without a mesoporous SiO2 shell broke HDPE chains randomly to-
ward alkane products with wide carbon chain distributions and lower yields (52.8%).
Selective hydrogenolysis is preceded by a processive mechanism that resembles
enzyme-catalyzed depolymerization of biomacromolecules to atom-precise frag-
ments. The long PE chains readily enter the nanopores and are then bound at a
desired position via polymer-surface interactions. Subsequently, the Pt nanopar-
ticles selectively cleave the C–C bond in the PE at regular intervals. Finally, small-
molecule products are desorbed and released because of the weak interaction
between the small molecules and the catalyst.

As described in the section of other routes to useful chemicals, tandem catalysis has
provided an effective strategy for the upcycling of PE and PET to value-added chem-
icals without consuming external hydrogen. Tandem catalytic cross-alkane metath-
esis (CAM) using low-value and widely available short alkanes (Figure 5) as cross-
metathesis partners enables the efficient conversion of different types of PE to liquid
fuels and waxes under relatively mild conditions (175 C).40 During the reaction, the
iridium (Ir) complex catalyzes the dehydrogenation of both PE and short alkane,

Cell Reports Physical Science 2, 100514, August 18, 2021 13


ll
OPEN ACCESS
Review

Figure 5. Tandem catalytic cross alkane metathesis of PE


Tandem catalytic cross-alkane metathesis of PE with short alkanes toward liquid fuels and waxes.40

generating unsaturated alkenes and Ir–H2. Next, the Re2O7/Al2O3 catalyst breaks
down the PE chains by scrambling the double bond. Finally, Ir–H2 hydrogenates
the newly formed alkenes to saturated alkanes. PE fragments with long chains can
react with the light alkane via multiple CAM processes, eventually leading to short
hydrocarbons.

The combination of Pt/WO3/ZrO2 and HY zeolites also enables efficient conversion


of polyolefins into branched, liquid fuels, including diesel, jet, and gasoline-range
hydrocarbons via a tandem catalysis process: polymers are first converted into rela-
tively large olefins or alkanes via hydrocracking over Pt/WO3/ZrO2; subsequently,
the intermediates diffuse into HY zeolite and then crack to form smaller alkenes.95
Because of the synergistic effect of the two catalysts, high yields (85%) of fuel prod-
ucts are obtained under relatively mild temperature (225 C, 3 MPa H2).

Plastic wastes can also be valorized to more-valuable fuel products, such as jet fuel,
with the rapid development of accurate and selective depolymerization and hydro-
genolysis technologies. For example, hydrogenolysis of HDPE over a Ru/C catalyst
can provide jet-fuel and lubricant-range hydrocarbons with yields as high as 60.8 and
31.6 wt%, respectively.96 Theoretically, the chemical composition of plastic wastes is
close to that of oxygen-containing liquid fuels, which have distinct advantages over
current alkane fuels. Compared with lignocellulosic biomass, the upcycling of plastic
wastes to oxygen-containing liquid fuels will consume less, or even zero, external
hydrogen if desired reaction pathway can be reached.

UPCYCLING TO MATERIALS
Carbon materials
As abundant and low-cost carbon sources, plastic wastes have been used extensively
to produce carbon materials.97 The conversion of plastic wastes to versatile carbon
materials could not only expand in applications, but could also acquire the benefit
of carbon sequestration to some degree. Thanks to the richness and variety of plastic
waste and the processing techniques, multifarious carbon materials, including
carbon dots (C-dots),98 carbon microspheres,99 carbon nanofibers,100 carbon

14 Cell Reports Physical Science 2, 100514, August 18, 2021


ll
OPEN ACCESS
Review

nanosheets (CNSs),101 three-dimensional (3D) porous carbon,102 CNTs,49,103–105


graphite,106 and graphene107,108 have been manufactured from plastic wastes.

Conversion of plastic wastes to value-added CNTs and graphene, which show


tremendous prospects in many domains of science and engineering, represents
two important directions of upcycling. Conversion of polyolefin plastics, including
PE and PP, to CNTs is usually achieved via two-step process.109,110 Polyolefin plas-
tics are first converted to gas products via pyrolysis, and then the gas products are
transformed to CNTs over Ni or Fe-based catalysts, with yields limited by the effi-
ciency of both pyrolysis and catalytic CNTs growth process.103 Recently, as
described in the section of upcycling to fuels, the microwave-initiated catalytic
deconstruction over an FeAlOx catalyst has enabled a one-pot conversion of PE,
PP, and PS to MWCNTs without undergoing gas intermediates, providing a simple,
rapid, and efficient CNT-production process. A flash Joule heating (FJH) process en-
ables gram-scale, bottom-up synthesis of high-quality flash graphene using mixed
plastic waste composed of PET, HDPE, LDPE, PVC, PP, and polyacrylonitrile as the
carbon source.108 This process delivers turbostratic graphene with an interlayer
spacing of 3.45 Å, with carbon oligomers, hydrogen, and light hydrocarbons as
by-products.111 The upcycling of plastic waste to graphene is economically attrac-
tive for scale-up because of its low energy consumption ($125 in electricity per
ton of plastic waste), which is even less than that of producing virgin plastic. Because
the FJH process works for various carbon sources, including mixed plastic waste, dis-
carded food, and rubber tires, without the use of a furnace, solvents, or reactive
gases, it is a promising method for treating landfill plastic waste.

Plastic-waste-derived carbon materials are promising candidates for adsorbents,112


electrocatalysts,113 supercapacitors,114 batteries.102 and solar vapor generation ma-
terials.115 For instance, the PE based CNSs are successfully used as transparent con-
ducting film (TCF) to fabricate organic photovoltaic cells.101 The use of PP-derived
CNS which has thin thickness (4-4.5 nm) and large specific surface area (3200 m2
g 1) with well-defined hierarchical porous structure as symmetric supercapacitor ex-
hibits high energy density.116 The sulfonated carbon scaffold, which is obtained by
microwave-promoted sulfonation of LDPE followed by pyrolysis at 900 C, can be
employed as interlayer in high-performance lithium-sulfur batteries.117

Polymers
The monomers derived from plastic depolymerization are usually returned to the
manufacturer of the original plastic. In addition, the monomers and their derivates
made through further chemical or enzymatic transformation can be used for produc-
tion of new materials. For example, TPA can be converted to 2-pyrone-4,6-dicarbox-
ylic acid, a promising building block for bio-degradable plastic, by two recombinant
Escherichia coli strains.118 Another option is to incorporate plastic-derived mono-
mers, oligomers, or even polymer fragments into new materials through copolymer-
ization with external building blocks. One-pot upcycling of BPA-PC into high-value
poly(aryl ether sulfone)s (PSUs) (Figure 6A), a type of high-performance engineering
thermoplastics that can be used for reverse-osmosis and water-purification mem-
branes, medical equipment, and high-temperature applications, has been achieved
by depolymerization of BPA-PC into active phenoxides, which are then in situ poly-
condensed with daryl fluorides.119

Aminolysis of PET and BPA-PC provides multifarious modular scaffolds for new poly-
mer production. PET can be transformed to poly(aryl ether sulfone-amide), an
advanced thermoplastic material, via aminolysis with 4-aminophenol, followed by

Cell Reports Physical Science 2, 100514, August 18, 2021 15


ll
OPEN ACCESS
Review

Figure 6. Incorporation of partial monomers into new polymers


(A) One-step conversion of BPA-PC into poly(aryl ether sulfone)s. 119
(B) Conversion of PET to poly(aryl ether sulfone-amide). 120
(C and D) Aminolysis of PET (C) and BPA-PC (D) with imidazole-amines to bis-imidazole monomers. 121
(E) Diaminolysis of BPA-PC wastes with long-chain diamines toward non-isocyanate polyureas. 122
(F) Conversion of BPA-PC to aliphatic polycarbonates.123 TBD:MSA, 1,5,7-triazabicyclo[4.4.0]dec-5-ene:methane sulfonic acid; DMAP, N,N-dimethyl-4-
aminopyridine; CHP, N-cyclohexyl-2-pyrrolidone.

nucleophilic aromatic substitution polymerization (Figure 6B).120 Aminolysis of PET


and BPA-PC with imidazole-amines results in EGs with amide or urea functionality-
containing monomers (Figure 6C and Figure 6D), providing an alternative process
to produce bis-imidazole monomers without the use of toxic chloride reagents.121
The PET-derived bis-imidazole monomers can be used to produce high-

16 Cell Reports Physical Science 2, 100514, August 18, 2021


ll
OPEN ACCESS
Review

Figure 7. Upcycling of plastic wastes to vitrimers


(A) Conversion of commercial HDPE to vitrimers via metathesis with dioxaborolane. 130 Adapted with permissions, 130 copyright 2017, AAAS.
(B) Conversion of PP to vitrimers by grafting MA. 131

molecular-weight ionic polymers, a type of high-performance material with good


elastomeric and self-healing property. Similarly, aminolysis of PET with amino-alco-
hols affords diol terephthalamides, which can be used to synthesize poly(ester-
amide)s for new polymer production.124 Depolymerization of PET with o-phenylene-
diamine and 2-aminophenol provides bis-benzimidazole and bis-benzoxazole,
respectively, which are promising feedstocks for medicines, high-performance poly-
mers, and organic electronics.125 The diaminolysis of BPA-PC with long-chain di-
amines in THF produces bisphenol A and non-isocyanate polyureas (Figure 6E)
through a non-isocyanate and one-pot route.122 The depolymerization of BPA-PC
with aliphatic diols produces bisphenol A and the corresponding carbonate-contain-
ing diols (Figure 6F), which can be polymerized into a series of value-added aliphatic
polycarbonates with ionic conductivity as potential polymer electrolytes for solid-
state batteries.123

The development of covalent adaptable networks that can transit their behavior un-
der specific stimuli has long been of interest to researchers as a promising approach
toward bridging the gap between thermoplastics and thermosets.126,127 Upcycling
of post-consumer, plastic waste into covalent adaptable networks could impart re-
cyclability and retain many of the beneficial properties of the thermoset materials.
Vitrimers are one typical type of a covalent, adaptable network that combines the
advantages of thermoplastics (light, tough, and easy to process, assemble, and
recycle through heating) and thermosets (high mechanical, thermal, and chemical
resistance).126,128,129 Because vitrimers are composed of polymer networks that un-
dergo rearrangement of chemical bonds and topology via, for example, exchange
reactions, without reducing the number of chemical bonds and cross-links, they
can reversibly switch between thermosets and thermoplastics.130 Commercial
HDPE, a typical thermoplastic, has been converted to high-performance vitrimers
through grafting of maleimides bearing dioxaborolane functionalities onto HDPE,
followed by cross-linking of functional polymers containing pendant dioxaborolane
units, via metathesis with a bis-dioxaborolane (Figure 7A).130 Compared with the
original HDPE, the resulting vitrimers exhibit markedly improved melt strength,

Cell Reports Physical Science 2, 100514, August 18, 2021 17


ll
OPEN ACCESS
Review

dimensional stability at elevated temperatures, and solvent and environmental


stress-cracking resistance. Moreover, these vitrimers can be extruded or injection-
molded, akin to thermoplastics, which is a prominent advantage over thermosets
and vitrimers produced by traditional cross-linking methods. The dioxaborolane
metathesis method also enables the direct processing of incompatible polymers,
providing a promising approach for the valorization of plastic waste without sorting.
Analogously, conversion of polyolefins, including PP and PE, to high-performance
vitrimers (Figure 7B) has been achieved by grafting maleic anhydride (MA) using di-
cumyl peroxide (DCP) as a free-radical initiator, followed by epoxy-anhydride curing
with a di-functional epoxy (DGEBA).131 Introducing dibutyltin dilaurate could impart
malleability into commercial cross-linked PUR, a typical thermoset material, thus
enabling its reprocessing at high temperatures via dynamic carbamate-exchange re-
action.132 After further twin-screw extrusion, the PUR filaments or films obtained
show elastomeric or rigid thermoset mechanical properties. Functionalization of
polyisoprene and PS by azidoformate grafting agents though reactive processing af-
fords vitrimers with good creep resistance at room temperature and good mechan-
ical properties.133 Incorporating dioxaborolane during the polymerization of methyl
methacrylate130 or treating vinyl-monomer-derived prepolymers with a trifunctional
amine134 could produce PMMA vitrimers, but direct upcycling of post-consumer
PMMA into vitrimers has not, to our knowledge, been achieved.

Post-functionalization is a potential strategy to impart certain functionalities and


unique properties to plastics and to preserve the integrity of the polymer skel-
eton.135,136 For instance, TBD-catalyzed or metal-catalyzed transesterification of
acrylic polymers, such as poly(methyl acrylate) (PMA) and PMMA in the presence
of alcohol and amine nucleophiles, combined with reversible-addition fragmenta-
tion chain transfer, allows for the precise functionalization of sterically differentiated
acrylic copolymers and polymeric chain ends, providing a new route for upcycling of
adhesive and coating waste into value-added materials.137,138 In contrast,
controlled defunctionalization may also afford unexpected materials. For example,
PMA can be converted to PP via B(C6F5)3-catalyzed deoxygenation in the presence
of silane.139 In addition, grafting of acrylic acid, poly(propylene imine) dendrimer,100
and acrylonitrile140 onto PP could produce functionalized adsorbents.

The integration of plastic valorization with biomass resources has provided an


effective strategy to dramatically improve economic and environmental benefits.
As a proof-of-concept, the combination of partially glycolized PET with bio-based
monomers enables the upcycling of PET waste into high-value composite materials
(Figure 8A), synchronously incentivizing plastic reclamation and the actual applica-
tion of biomass-derived chemicals.39 PET is partially glycolized with a diol (EG or
1,4-butanediol) over a titanium-butoxide catalyst. Afterward, glycolized PET is con-
verted to unsaturated polyesters (UPEs). Fiberglass-reinforced plastics (FRPs) are
produced by dissolving the UPEs in a reactive diluent in the presence of a free-
radical initiator. Alternatively, FRPs can also be directly synthesized from partially
deconstructed PET with olefinic carboxylic acid via polymerization followed by
cross-coupling by free-radical initiator. Compared with petroleum-based FRPs,
the two types of FRPs prepared from PET show substantially improved material
properties with a total supply-chain energy savings of 57% and GHG emission
reduction of 40%. Moreover, PET-derived FRPs have a higher selling price
($1.85 USD/lb) than PET resin does ($0.81 USD/lb) with 75% lower energy con-
sumption than that of returning the product to bottle-grade PET resin. Similarly,
transesterification of poly(glycolic acid) (PGA with bioderived g-butyrolactone
[BL] as both the comonomer and the solvent yields a sequence-defined copolymer

18 Cell Reports Physical Science 2, 100514, August 18, 2021


ll
OPEN ACCESS
Review

Figure 8. Incorporation of plastic fragments into new materials


(A) Upcycling of PET to FRPs via partial glycolysis, followed by combination with biomass-derived monomers. 39
(B) Transesterification of PGA with g-butyrolactone into a copolymer. 141
(C and D) Controlled isoselective polymerization of propylene and ethylene toward multiblock copolymers over isoselective pyridylamidohafnium
catalyst (iPr, isopropyl) activated by B(C 6 F 5 ) 3 (C) and the use of copolymers to compatibilize commercial PE and iPP (D). 142 (D) was adapted with
permissions, 142 copyright 2017, AAAS.

poly(GA-co-BL)) (Figure 8B), which is not only chemically recyclable but also ex-
hibits improved thermal stability (R44 C) than PGA.141

The use of plastic waste as the feedstock in additive manufacturing (AM) maps out a new
path for plastic recycling and upcycling toward a circular economy.143 For instance,

Cell Reports Physical Science 2, 100514, August 18, 2021 19


ll
OPEN ACCESS
Review

blends of HDPE waste with low-molecular-weight PP and MA-modified PP in a specified


formula can be converted into filaments for 3D printing with properties comparable to
PLA, one of the most widely used thermoplastics for AM.144 In another example, the
waste from children’s toys, which are composed of acrylonitrile butadiene styrene,
were successfully converted to filaments comparable to virgin materials.145

Composite materials
Current plastic-recycling technologies require stringent sorting and separation
steps to recover pure polymer as the feedstock because different types of polymers
in the actual plastic waste are usually immiscible.27,146 Because of the formation of
separate phases with weak mechanical strength at the interfaces, the melted blends
are brittle, even when just containing a small quantity of the other type of poly-
mer.147 The use of compatibilizers as molecular stitches to anchor different polymers
allows for direct upcycling of mixed plastic waste toward high-value materials,
exempting the costly and time-intensive sorting step.27,147–149 For example, a series
of semicrystalline PE/iPP multiblock copolymers, which can be prepared by
controlled isoselective polymerization of propylene and ethylene (Figure 8C), ex-
hibited superior efficiency for the compatibilization of PE and iPP.142 PE-block-iPP
diblock copolymers (Figure 8D) can weld common grades of commercial PE and
iPP together only when the molecular weight of the blocks exceeds a threshold. In
contrast, just 1% of tetrablock copolymer could transform mixed polyolefin munic-
ipal waste (typically PE:iPP = 70:30) into mechanically tough blends with consider-
ably higher strain at the break point (εb = 450%) than that of direct blending (εb =
12%) because of the formation of entangled loops, which can stitch the homopoly-
mers together effectively and form block copolymer films upon crystallization.
Analogously, the multiblock copolymers consisting of PP and polar PU segments
can be used as a versatile compatibilizer for PP and polar polymers, including
poly(methyl methacrylate), PVC, and poly(vinylidene difluoride).150 Poly(ethylene
terephthalate)-polyethylene multiblock copolymers also serve as both the adhe-
sive-promoting layer and the compatibilizer to recycle PET/PE mixed waste.151
With MA, a commercial compatibilizer, recycled PP and PET blends can be remolded
to water pipes with improved flexural properties than that of virgin PP without
compromising its impact properties.152 In addition, block copolymers prepared by
controlled polymerization could be used to prepare vitrimers with considerably bet-
ter resistance to macroscopic deformation than their analogs derived from statistical
copolymers.153 Therefore, in the future, it may be possible to convert plastic waste to
high-performance vitrimers via block copolymers as intermediates.

In addition to the blending of different types of polymers, plastic waste can also be
recombined with other types of materials toward the manufacture of composite
and hybrid materials. The use of plastic waste to displace virgin plastic offers an effec-
tive way to reduce the cost of fabricating commercial composite materials. For
example, PET/cotton blend fabrics are widely used in garments and home textiles,
accounting for more than 50% of textiles.154 Polyester waste can be used as the feed-
stock to produce cotton-based composite fabric with improved mechanical proper-
ties.155 Polyester/cotton blend fabrics can be further converted to a fiber-reinforced
composite through co-depositing of polyethylenimine and catechol to enhance the
interfacial bonding of the fiber matrix.154 Pseudopolyrotaxane inclusion complex
with improved thermal stability and optical properties can be prepared from styro-
foam (also called expanded polystyrene), by sulphonation, followed by encapsula-
tion of g-cyclodextrin.156 Both PET and PLA can be used as inker sources to synthesize
value-added metal-organic frameworks (MOFs) through either one-pot or two-step
processes, using the corresponding monomers as linkers.157,158 The use of PET waste

20 Cell Reports Physical Science 2, 100514, August 18, 2021


ll
OPEN ACCESS
Review

A
Hydrothermal hydrolysis (120 C) Degradation of PVC by Electro-Fenton-
Abiotic degradation combined with catalytic oxidation of PET like system using TiO2/graphite as
and biodegradation by peroxymonosulfate over magnetic N- cathode under heating (100 C) :
slow. doped nanocarbon springs: TOC dechlorination efficiency of 75 % with
removal rate of 44%. weight loss of 56%.

Limited photocatalytic degradation Photocatalytic degradation of PS Photocatalytic mineralization of PE,


of PE over ZnO-Pt, N-TiO2 , BiOCl and PE over TiO2 nanoparticle PP, PVC over single‐unit‐cell thick
morphology changes, appearance of film under 254 nm UV light: Nb2O5 layers under simulated
some oxygen groups and increase of disappearance of plastic environment conditions: 100%
carbonyl index. fragment in 24 and 36 h. degradation in 40, 60, 90 h.

B C
O2 Photoreduction H2O Photooxidation O2
Simulated environment 100% degradation
PE
conditions: 100 mW/cm2, CO2 (predominant); O2·- H2O2
H2O, air, 25 C
trace-level CO; ·OH + O2 ·COOH HOOC-COOH
PP
Single-unit-cell thick CH3COOH (0.56, 0.49, Fast e- + H+ e- + H+
Slow
Nb2O5 layers 0.24 μg/gplatic-1 at 36 h) CO2 CH3COOH
PE C-C bond cleavage C-C coupling
PVC

Figure 9. Research highlights in catalytic degradation of plastic wastes


(A) Trend of the development of catalytic degradation technologies.
(B and C) Degradation of PE, PP, and PVC under simulated natural environment conditions using single-unit-cell-thick Nb 2 O 5 layers as photocatalyst (B)
and the proposed reaction mechanism (C). 65

as the source and support to grow UiO-66 MOF provides hybrid PET@UiO-66 mate-
rial.159 As a novel sorbent, PET@UiO-66 exhibits obvious advantages, such as high
adsorption capacity and rate, stability and recyclability and 100 times enhanced
permeability than UiO-66 powder.

CATALYTIC DEGRADATION
Advanced oxidation processes (AOPs), including Fenton reactions, Fenton-like reac-
tions, and photocatalytic oxidation have offered a potential portfolio for the catalytic
degradation of plastic waste (Figure 9A). These technologies were primarily de-
signed and developed for the degradation of soluble organic pollutants with low
molecular weights. Previously, AOPs have been widely investigated to degrade bi-
sphenols and phthalates plasticizers, such as bisphenol A and diethyl phthalate.160 A
few hydrolyzable plastics can be depolymerized to monomers and then be further
treated in the form of molecules by AOPs. For instance, the degradation of polyes-
ters can be achieved by depolymerizing ester bonds by 80% hydrazine hydrate along
with NaOH, followed by the Fenton reaction.161 Nevertheless, most plastics are
insoluble macromolecules, which cannot be depolymerized under normal condi-
tions. To achieve efficient degradation of plastic waste, AOPs must be improved
or combined with other means, according to their structural features.

AOPs and hydrothermal hydrolysis have been successfully integrated to degrade


PE-based microplastics into environmentally benign organic intermediates in

Cell Reports Physical Science 2, 100514, August 18, 2021 21


ll
OPEN ACCESS
Review

water.64 The magnetic N-doped nanocarbon springs (Mn@NCNTs) catalyst, which is


prepared by encapsulating manganese carbide nanoparticles into helically N-doped
carbon nanotubes, provides excellent performance for the degradation of micro-
plastics (size ranges from 0.01 to 1.5 mm) from commercial facial cleanser by acti-
vating peroxymonosulfate to generate sulfate radicals and hydroxyl radicals ($OH).
Microplastic removal efficiency (weight loss) of 44% was achieved at 120 C for 8 h.
The degradation products are mainly composed of low-molecular-weight and
less-toxic organic compounds, which can be used as a carbon source for algae culti-
vation. The removal efficiency declines by 45% in the fifth consecutive cycle.

Electro-Fenton-like systems could promote the degradation of PVC microplastics via


the cooperation of reduction and oxidation reactions. Using TiO2/graphite as the
cathode, the potentiostatic electrolysis at 0.7 V versus Ag/AgCl under heat
(100 C) for 6 h attains a dechlorination efficiency of 75% with a PVC weight loss of
56%.162 The dechlorination of PVC is achieved through direct reduction by the
applied cathode potential, whereas the cleavage of the PVC backbone is dominated
by $OH.

In addition to Fenton and Fenton-like reactions, photocatalytic oxidation have been


studied extensively for catalytic degradation of plastic waste. Under visible light,
degradation of PE microplastics (1 mm 3 1 mm) over a plasmonic platinum/zinc ox-
ide nanorod (ZnO-Pt) catalyst163 and degradation of LDPE-bag debris (in the form of
films with sizes from 3 mm 3 3 mm to 5 mm 3 5 mm) over a mesoporous N-TiO2 cata-
lyst164 in an aqueous solution only leads to a slight deconstruction of plastic,
including morphology changes, the appearance of some oxygen groups, and an in-
crease of carbonyl index, without reporting weight loss. The degradation of HDPE
microspheres (200–250 mm) in an aqueous solution over a hydroxy-rich, ultrathin
BiOCl catalyst under simulated solar light for 5 h attains a weight loss of 5.38%.165
Comparatively, the degradation of smaller plastic fragments is much easier than
the larger ones.164,165 Recently, efficient degradation of small-size PS microspheres
(5 mm) and PE (approximately 50 3 50 mm) to CO2 via solid-phase photodegradation
was achieved over a TiO2 nanoparticle film prepared with Triton X-100,15 but the
degradation efficiency was evaluated by the reduction in volume of the PS sphere,
which was estimated by scanning electron microscopy (SEM) analysis, instead of
the more-precise weight analysis. Complete disappearance of PS and PE in SEM
images was achieved under 254 nm ultraviolet (UV) light within 24 and 36 h, respec-
tively. The degradation efficiency of microplastic at the solid-solid interface (micro-
plastic-TiO2) is considerably higher than that in the liquid phase because the reactive
species can attack the polymers directly without any hindrance through the direct
contact of microplastic with the TiO2 film.

Complete mineralization of PE, PP, and PVC into CO2 (Figure 9B) has been achieved
via photocatalytic degradation using single-unit-cell-thick Nb2O5 layers as photoca-
talyst under simulated natural environment conditions (Xe lamp with a standard AM
1.5G filter, 100 mW/cm2, H2O, air, room temperature) within 40, 60, and 90 h,
respectively.65 The 100% mineralization was demonstrated by the vanished charac-
teristic peaks of the plastics in X-ray diffraction (XRD) and Raman spectra, as well as
by the formation of CO2 with total moles close to the C content in the feedstock. The
100% mineralization under simulated environment conditions is inspiring—which
has not been reached by other AOPs. Because the valence band (+2.5 V versus
normal the hydrogen electrode [NHE; pH 7]) of Nb2O5 is higher than the redox po-
tential of H2O/$OH (+2.32 V NHE) and the conduction band (0.9 V versus NHE) is
lower than the required potentials of CO2 reduction (0.6 V versus NHE, pH = 7),

22 Cell Reports Physical Science 2, 100514, August 18, 2021


ll
OPEN ACCESS
Review

the Nb2O5 catalyst allows a one-pot conversion of the plastics to acetic acid
(CH3COOH) through a sequential, photoinduced C–C bond cleavage and coupling
pathway. Along with the photodegradation of the plastic toward CO2 through
oxidative C–C cleavage triggered by O2 and $OH radicals (Figure 9C), the generated
CO2 is further selectively photoreduced by H+ and e toward CH3COOH, a potential
high-energy-density C2 fuels. Because of the low rate of photoreduction, the
CH3COOH yields after 36 h were just 0.56, 0.49, and 0.24 mg/gplatic 1 for PE, PP,
and PVC, respectively. Similarly, vanadium photocatalysts have been reported to
be capable of catalyzing the cascade carbon-carbon-bond scission of a series of hy-
droxyl-terminated polymers, including polyethylene glycol and polyethylene mono-
alcohol (a model of PE) under visible light irradiation without alkaline pretreatment,
but their catalytic performance for actual plastic waste was not provided.166

Although noteworthy progress have been achieved, the application of AOPs to


remediate microplastic pollution in actual wastewater and natural aquatic environ-
ments faces considerable challenges. Except for the Nb2O5-based photocatalytic
system, the degradation of microplastics in most catalytic systems is very slow at
ambient temperature, whereas heating actual wastewater and natural aquatic envi-
ronments to the evaluated temperatures requires a large quantity of energy. The
coexistence of microplastics with biodegradable organic matter may consume large
amounts of oxidizing agents, increasing the cost considerably. Furthermore, the
reuse of catalysts seems to be impossible because of the difficulties in catalyst recov-
ery. Therefore, much effort is required to develop more-powerful and selective
degradation technologies.

CONCLUSIONS AND FUTURE PROSPECTS


Although there is no panacea for the complicated and successionally evolutive plas-
tics problem, upcycling and catalytic degradation map out two important routes to-
ward the vision of a sustainable and zero-pollution future. The concept of upcycling
was developed to ameliorate the intrinsic drawbacks of traditional recycling technol-
ogies and to improve the overall benefits through unearthing of products of high
value. The ready availability and diversity of plastic waste represent a largely un-
der-developed opportunity. Tailored upcycling processes for each type of plastic
are the most likely approach to deliver maximal economic and environmental bene-
fits.31,57 Compared with conventional recycling processes and many industrialized
processes based on fossil resources, some upcycling processes and products have
already been shown to have distinct advantages in both high-value areas and basic
applications. Ascertaining the embedded value in plastic waste, which is present in
the form of carbon, hydrogen, chemicals, energy, and macromolecular structures,
might not only reduce the generation of undesired side-products and the consump-
tion of energy and other external resources but also contribute to the improvement of
conversion efficiency, eventually elevating the overall benefits. Furthermore, some
upcycling processes have shown superiority in the handling of real-world, mixed,
or contaminated plastic waste with a decrease in process complexity and cost. Impor-
tantly, most upcycling processes are compatible with both existing sorting and recy-
cling processes, and the feasibility of using or retrofitting existing infrastructure might
facilitate an uptake in new routes. The lucrative upcycling processes would not only
enhance the reclaiming rate but also, in turn, expedite the development and actual
application of traditional recycling processes. Indeed, there are also some controver-
sial upcycling processes that involve the use of toxic reagents or result in additional
issues, such as secondary pollution. Notably, the resulting materials do not neces-
sarily mean they can be subjected to a continuous recycling and upcycling process.

Cell Reports Physical Science 2, 100514, August 18, 2021 23


ll
OPEN ACCESS
Review

Conferring degradability or recyclability into the target materials during the upcy-
cling process is significant toward establishing an ideal, sustainable, closed loop.

Upcycling technologies can be enhanced by the following endeavors. Advanced cata-


lytic methods must be established to improve reaction efficiency. Generally, both upcy-
cling and recycling processes are restrained by limited solid-solid contact efficiency
between the recalcitrant plastic substrate and heterogenous catalysts, intricate reaction
networks, and undesired degradation of the target products. Hence, the development
of quantum dot catalysts, homogeneous catalysts, enzyme-mimetic catalysts, tandem
catalysis strategies, stabilization chemistry, and the coupling of different catalytic tech-
nologies is a potential option for solving these bottlenecks. In addition, the catalysts
should have a high tolerance for air, moisture, and organic or metal salt contaminants.37
Beyond catalysts, the screening of applicable solvents able to swell, dissolve, or selec-
tively separate different polymers167 or able to remove colorants and other additives is
also essential.31 A comprehensive upcycling roadmap should be established and then
be combined with recycling processes on the cusp of commercialization to make them
available for most plastic waste streams. It should be emphasized that, under the frame-
work of upcycling, plastic waste can be considered an important and unique resource
that ranges between fossil and biomass resources, not solely a troubling litter. There-
fore, the development of plastic upcycling should learn from the well-established oil
refining processes and rapidly growing biorefinery processes. The drop-in integration
of plastic upcycling with oil refining, biorefinery, and CO2 use processes would foster
new reaction routes and products. In fact, similar chemical-bond activation strate-
gies,168,169 enzymatic decomposition mechanisms,29 and photoreforming process86
have already been successfully applied for the conversion of both biomass and plastic
waste. In addition, one-size-fit-all upcycling approaches for mixed plastic waste should
be explored to obviate the need for sorting, separating, and purifying.27 Whether up-
cycling processes are sustainable and technically, economically, and environmentally
viable depends on the entire life cycle from collection and sorting of plastic waste to
production of high-value products to end-of-life management. Therefore, the eco-
nomic benefits and environmental impacts must be evaluated on a case-by-case basis
with rigid criteria and be compared with appropriate benchmarks before their
commercialization.

The development of degradation technologies is still at its early stages. Up until now,
most degradation methods developed for molecular pollutants have been substanti-
ated to be ineffective, whereas a few of catalytic processes have already showed the
possibility of degrading plastic waste under artificially optimized harsh conditions
with a series of ancillary measures and superiority over biological degradation. More
scientific proof, based on tracing experiments or monitoring of degradation products
using mass balances, is required to confirm the claimed degradation and mineraliza-
tion.33 Achieving efficient degradation of real-world plastic waste under actual environ-
mental conditions remains an important and challenging task for future research. The
aforementioned strategies aimed at improving the upcycling processes should be help-
ful for enhancing catalytic degradation technologies; among which, the selective
contact and binding of substrates with catalysts or the screening of cheap and environ-
mentally friendly catalysts and oxidants is of great importance. Compared with miner-
alization, partial and selective degradation toward environmentally benign products or
degradable products, followed by spontaneous biodegradation seems to be a more-
viable option for reducing the consumption of energy and oxidants.

The ongoing modification of currently used plastic materials and the burgeoning in-
vention of new polymers create both opportunities and challenges for plastic

24 Cell Reports Physical Science 2, 100514, August 18, 2021


ll
OPEN ACCESS
Review

recycling, upcycling, and degradation. Meanwhile, the use-pattern and scale of plas-
tic materials are also varying rapidly.170 Many plastic packaging materials manufac-
tured in modern industry are multilayer plastic films composed of distinct polymer
layers.8,167 The complex composition and tough structure considerably increase
the difficulty of sorting and reprocessing, impeding the economic benefits of recy-
cling and upcycling. The design of recyclable plastics that undergo reversible (de)
polymerization171,172 and recyclable thermosetting plastics containing dynamic co-
valent bonds173–175 would not only promote chemical and mechanical recycling but
also influence upcycling processes. In addition, the possible contamination of plastic
wastes by the decomposition of degradable plastics may also increase the separa-
tion costs and reduce the performance of recovered materials.35

Another pivotal environmental concern associated with plastics is the exhaustion of


non-renewable fossil resources. The most hopeful solution to this problem is to
develop more ‘‘sustainable’’ polymers from CO2 and renewable resources, in partic-
ular, plant-based feedstocks, such as cellulose, lignin, other polysaccharides, deox-
yribonucleic acid, polyhydroxyalkanoates, sugars, vegetable oils, and terpenes;
however, most current routes toward renewable feedstocks are far from cost effec-
tive relative to conventional routes.2,34,176–183 Furthermore, sustainable polymers
from renewable resources do not mean they are also degradable or recyclable.
Although several recyclable and degradable polymers from renewable sources
have been successfully designed in the laboratory,178,184–188 the ideal integration
of sustainability, degradability (or recyclability), usability, and low-cost, in practice,
is still a huge challenge. Therefore, sustainable polymers also require the corre-
sponding recycling, upcycling, and degradation technologies once they are pro-
duced on a large scale.

ACKNOWLEDGMENTS
This work was supported by the National Natural Science Foundation of China
(21878163), the China Postdoctoral Science Foundation (2018M640231), the Funda-
mental Research Funds for the Central Universities, National Key Research Project
(2018YFD080083-03), the 2017 Science and Technology Demonstration Project of
Industrial Integration and Development, Tianjin, China (17ZXYENC00100), the Na-
tional Natural Science Foundation of China (51708301), and the Young Elite Scien-
tists Sponsorship Program by Tianjin (TJSQNTJ-2018-06).

AUTHOR CONTRIBUTIONS
Q.H. proposed the topic of the review and wrote the manuscript. M.Z., H.Q., Y.N.,
X.B., T.X., M.L., and Q.L. investigated the literature. M.J. revised a draft of the
review.

DECLARATION OF INTERESTS
There are no conflicts of interest to declare.

REFERENCES
1. Sardon, H., and Dove, A.P. (2018). Plastics 4. Brooks, A.L., Wang, S., and Jambeck, J.R. 7. Arp, H.P.H., Kühnel, D., Rummel, C.,
recycling with a difference. Science 360, (2018). The Chinese import ban and its impact MacLeod, M., Potthoff, A., Reichelt, S., Rojo-
380–381. on global plastic waste trade. Sci. Adv. 4, Nieto, E., Schmitt-Jansen, M., Sonnenberg,
eaat0131. J., Toorman, E., and Jahnke, A. (2021).
2. Hillmyer, M.A. (2017). The promise of plastics Weathering plastics as a planetary boundary
from plants. Science 358, 868–870. 5. Gibb, B.C. (2019). Plastics are forever. Nat. threat: exposure, fate, and hazards. Environ.
Chem. 11, 394–395. Sci. Technol. 55, 7246–7255.
3. Geyer, R., Jambeck, J.R., and Law, K.L. (2017).
Production, use, and fate of all plastics ever 6. Horejs, C. (2020). Solutions to plastic 8. Lau, W.W.Y., Shiran, Y., Bailey, R.M., Cook, E.,
made. Sci. Adv. 3, e1700782. pollution. Nat. Rev. Mater. 5, 641. Stuchtey, M.R., Koskella, J., Velis, C.A.,

Cell Reports Physical Science 2, 100514, August 18, 2021 25


ll
OPEN ACCESS
Review

Godfrey, L., Boucher, J., Murphy, M.B., et al. 21. Oliveri Conti, G., Ferrante, M., Banni, M., (2020). Chemically triggered synthesis,
(2020). Evaluating scenarios toward zero Favara, C., Nicolosi, I., Cristaldi, A., Fiore, M., remodeling, and degradation of soft
plastic pollution. Science 369, 1455–1461. and Zuccarello, P. (2020). Micro- and nano- materials. J. Am. Chem. Soc. 142, 3913–3922.
plastics in edible fruit and vegetables. The
9. Martin, C., Baalkhuyur, F., Valluzzi, L., first diet risks assessment for the general 37. Garcı́a, J.M. (2016). Catalyst: design
Saderne, V., Cusack, M., Almahasheer, H., population. Environ. Res. 187, 109677. challenges for the future of plastics recycling.
Krishnakumar, P.K., Rabaoui, L., Qurban, Chem 1, 813–815.
M.A., Arias-Ortiz, A., et al. (2020). Exponential 22. Barboza, L.G.A., Dick Vethaak, A., Lavorante,
increase of plastic burial in mangrove B.R.B.O., Lundebye, A.-K., and Guilhermino, 38. Tournier, V., Topham, C.M., Gilles, A., David,
sediments as a major plastic sink. Sci. Adv. 6, L. (2018). Marine microplastic debris: an B., Folgoas, C., Moya-Leclair, E., Kamionka,
eaaz5593. emerging issue for food security, food safety E., Desrousseaux, M.L., Texier, H., Gavalda,
and human health. Mar. Pollut. Bull. 133, S., et al. (2020). An engineered PET
10. Bergmann, M., Mützel, S., Primpke, S., 336–348. depolymerase to break down and recycle
Tekman, M.B., Trachsel, J., and Gerdts, G. plastic bottles. Nature 580, 216–219.
(2019). White and wonderful? Microplastics 23. Hernandez, L.M., Xu, E.G., Larsson, H.C.E.,
prevail in snow from the Alps to the Arctic. Sci. Tahara, R., Maisuria, V.B., and Tufenkji, N. 39. Rorrer, N.A., Nicholson, S., Carpenter, A.,
Adv. 5, eaax1157. (2019). Plastic teabags release billions of Biddy, M.J., Grundl, N.J., and Beckham, G.T.
microparticles and nanoparticles into tea. (2019). Combining reclaimed PET with bio-
11. Katija, K., Choy, C.A., Sherlock, R.E., Sherman, Environ. Sci. Technol. 53, 12300–12310. based monomers enables plastics upcycling.
A.D., and Robison, B.H. (2017). From the Joule 3, 1006–1027.
surface to the seafloor: How giant larvaceans 24. Cox, K.D., Covernton, G.A., Davies, H.L.,
transport microplastics into the deep sea. Sci. Dower, J.F., Juanes, F., and Dudas, S.E. 40. Jia, X., Qin, C., Friedberger, T., Guan, Z., and
Adv. 3, e1700715. (2019). Human consumption of microplastics. Huang, Z. (2016). Efficient and selective
Environ. Sci. Technol. 53, 7068–7074. degradation of polyethylenes into liquid fuels
12. Jambeck, J.R., Geyer, R., Wilcox, C., Siegler, and waxes under mild conditions. Sci. Adv. 2,
T.R., Perryman, M., Andrady, A., Narayan, R., 25. Ragusa, A., Svelato, A., Santacroce, C., e1501591.
and Law, K.L. (2015). Marine pollution: plastic Catalano, P., Notarstefano, V., Carnevali, O.,
waste inputs from land into the ocean. Papa, F., Rongioletti, M.C.A., Baiocco, F., 41. Martı́n, A.J., Mondelli, C., Jaydev, S.D., and
Science 347, 768–771. Draghi, S., et al. (2021). Plasticenta: first Pérez-Ramı́rez, J. (2021). Catalytic processing
evidence of microplastics in human placenta. of plastic waste on the rise. Chem 7, 1487–
13. Li, C., Busquets, R., and Campos, L.C. (2020). Environ. Int. 146, 106274. 1533.
Assessment of microplastics in freshwater
systems: A review. Sci. Total Environ. 707, 26. Zheng, J., and Suh, S. (2019). Strategies to 42. Ghisellini, P., and Ulgiati, S. (2020). Circular
135578. reduce the global carbon footprint of plastics. economy transition in Italy: achievements,
Nat. Clim. Chang. 9, 374–378. perspectives and constraints. J. Clean. Prod.
14. Koelmans, A.A., Besseling, E., and Shim, W.J.; 243, 118360.
Nanoplastics in the Aquatic Environment 27. Garcia, J.M., and Robertson, M.L. (2017). The
(2015). Critical Review. In Marine future of plastics recycling. Science 358, 43. Zhang, F., Zeng, M., Yappert, R.D., Sun, J.,
Anthropogenic Litter, M. Bergmann, L. 870–872. Lee, Y.-H., LaPointe, A.M., Peters, B., Abu-
Gutow, and M. Klages, eds. (Springer Omar, M.M., and Scott, S.L. (2020).
International Publishing), pp. 325–340. 28. Rahimi, A., and Garcı́a, J.M. (2017). Chemical Polyethylene upcycling to long-chain
recycling of waste plastics for new materials alkylaromatics by tandem hydrogenolysis/
15. Nabi, I., Bacha, A.-U.-R., Li, K., Cheng, H., production. Nat. Rev. Chem. 1, 0046. aromatization. Science 370, 437–441.
Wang, T., Liu, Y., Ajmal, S., Yang, Y., Feng, Y.,
and Zhang, L. (2020). Complete 29. Chen, C.-C., Dai, L., Ma, L., and Guo, R.-T. 44. Tennakoon, A., Wu, X., Paterson, A.L.,
photocatalytic mineralization of microplastic (2020). Enzymatic degradation of plant Patnaik, S., Pei, Y., LaPointe, A.M., Ammal,
on TiO2 nanoparticle film. iScience 23, biomass and synthetic polymers. Nat. Rev. S.C., Hackler, R.A., Heyden, A., Slowing, I.I.,
101326. Chem. 4, 114–126. et al. (2020). Catalytic upcycling of high-
density polyethylene via a processive
16. Sun, X.-D., Yuan, X.-Z., Jia, Y., Feng, L.-J., Zhu, 30. Coates, G.W., and Getzler, Y.D.Y.L. (2020).
Chemical recycling to monomer for an ideal, mechanism. Nat. Catal. 3, 893–901.
F.-P., Dong, S.-S., Liu, J., Kong, X., Tian, H.,
Duan, J.-L., et al. (2020). Differentially charged circular polymer economy. Nat. Rev. Mater. 5, 45. Haider, T.P., Völker, C., Kramm, J., Landfester,
nanoplastics demonstrate distinct 501–516. K., and Wurm, F.R. (2019). Plastics of the
accumulation in Arabidopsis thaliana. Nat. 31. Vollmer, I., Jenks, M.J.F., Roelands, M.C.P., future? the impact of biodegradable
Nanotechnol. 15, 755–760. White, R.J., van Harmelen, T., de Wild, P., van polymers on the environment and on society.
der Laan, G.P., Meirer, F., Keurentjes, J.T.F., Angew. Chem. Int. Ed. Engl. 58, 50–62.
17. Dawson, A.L., Kawaguchi, S., King, C.K.,
and Weckhuysen, B.M. (2020). Beyond 46. Kratish, Y., Li, J., Liu, S., Gao, Y., and Marks,
Townsend, K.A., King, R., Huston, W.M., and
mechanical recycling: giving new life to plastic T.J. (2020). Polyethylene terephthalate
Bengtson Nash, S.M. (2018). Turning
waste. Angew. Chem. Int. Ed. Engl. 59, 15402– deconstruction catalyzed by a carbon-
microplastics into nanoplastics through
15423. supported single-site molybdenum-dioxo
digestive fragmentation by Antarctic krill. Nat.
Commun. 9, 1001. 32. Jehanno, C., Pérez-Madrigal, M.M., complex. Angew. Chem. Int. Ed. Engl. 59,
Demarteau, J., Sardon, H., and Dove, A.P. 19857–19861.
18. Lu, H.-C., Ziajahromi, S., Neale, P.A., and
(2019). Organocatalysis for depolymerisation.
Leusch, F.D.L. (2021). A systematic review of 47. Cabanes, A., Valdés, F.J., and Fullana, A.
Polym. Chem. 10, 172–186.
freshwater microplastics in water and (2020). A review on VOCs from recycled
sediments: Recommendations for 33. Wei, R., Tiso, T., Bertling, J., O’Connor, K., plastics. Sustain. Mater. Technol. 25, e00179.
harmonisation to enhance future study Blank, L.M., and Bornscheuer, U.T. (2020).
comparisons. Sci. Total Environ. 781, 146693. 48. Lopez, G., Artetxe, M., Amutio, M., Alvarez, J.,
Possibilities and limitations of
Bilbao, J., and Olazar, M. (2018). Recent
biotechnological plastic degradation and
19. Wong, J.K.H., Lee, K.K., Tang, K.H.D., and advances in the gasification of waste plastics:
recycling. Nat. Catal. 3, 867–871.
Yap, P.S. (2020). Microplastics in the a critical overview. Renew. Sustain. Energy
freshwater and terrestrial environments: 34. Zhu, Y., Romain, C., and Williams, C.K. (2016). Rev. 82, 576–596.
prevalence, fates, impacts and sustainable Sustainable polymers from renewable
solutions. Sci. Total Environ. 719, 137512. 49. Jie, X., Li, W., Slocombe, D., Gao, Y.,
resources. Nature 540, 354–362.
Banerjee, I., Gonzalez-Cortes, S., Yao, B.,
20. Boelee, E., Geerling, G., van der Zaan, B., 35. Albertsson, A.-C., and Hakkarainen, M. (2017). AlMegren, H., Alshihri, S., Dilworth, J., et al.
Blauw, A., and Vethaak, A.D. (2019). Water Designed to degrade. Science 358, 872–873. (2020). Microwave-initiated catalytic
and health: from environmental pressures to deconstruction of plastic waste into hydrogen
integrated responses. Acta Trop. 193, 36. Sun, X., Chwatko, M., Lee, D.-H., Bachman, and high-value carbons. Nat. Catal. 3,
217–226. J.L., Reuther, J.F., Lynd, N.A., and Anslyn, E.V. 902–912.

26 Cell Reports Physical Science 2, 100514, August 18, 2021


ll
OPEN ACCESS
Review

50. Christensen, P.R., Scheuermann, A.M., (2020). Photocatalytic conversion of waste 77. DelRe, C., Jiang, Y., Kang, P., Kwon, J., Hall,
Loeffler, K.E., and Helms, B.A. (2019). Closed- plastics into C2 fuels under simulated natural A., Jayapurna, I., Ruan, Z., Ma, L., Zolkin, K., Li,
loop recycling of plastics enabled by dynamic environment conditions. Angew. Chem. Int. T., et al. (2021). Near-complete
covalent diketoenamine bonds. Nat. Chem. Ed. Engl. 59, 15497–15501. depolymerization of polyesters with nano-
11, 442–448. dispersed enzymes. Nature 592, 558–563.
66. George, N., and Kurian, T. (2014). Recent
51. Morris, J. (1996). Recycling versus developments in the chemical recycling of 78. Fernandes, A.C. (2021). Reductive
incineration: an energy conservation analysis. postconsumer poly(ethylene terephthalate) depolymerization of plastic waste catalyzed
J. Hazard. Mater. 47, 277–293. waste. Ind. Eng. Chem. Res. 53, 14185–14198. by Zn(OAc)2 $ 2H2O. ChemSusChem.
Published online March 5, 2021. https://doi.
52. Bernardo, C.A., Simões, C.L., and Pinto, 67. Raheem, A.B., Noor, Z.Z., Hassan, A., Abd org/10.1002/cssc.202100130.
L.M.C. (2016). Environmental and economic Hamid, M.K., Samsudin, S.A., and Sabeen,
life cycle analysis of plastic waste A.H. (2019). Current developments in 79. Guironnet, D., and Peters, B. (2020). Tandem
management options. A review. AIP Conf. chemical recycling of post-consumer catalysts for polyethylene upcycling: a simple
Proc. 1779, 140001. polyethylene terephthalate wastes for new kinetic model. J. Phys. Chem. A 124, 3935–
materials production: A review. J. Clean. 3942.
53. Borrelle, S.B., Ringma, J., Law, K.L., Prod. 225, 1052–1064.
Monnahan, C.C., Lebreton, L., McGivern, A., 80. Lu, S., Jing, Y., Feng, B., Guo, Y., Liu, X., and
Murphy, E., Jambeck, J., Leonard, G.H., 68. Cano, I., Martin, C., Fernandes, J.A., Lodge, Wang, Y. (2021). H2-free plastic conversion:
Hilleary, M.A., et al. (2020). Predicted growth R.W., Dupont, J., Casado-Carmona, F.A., converting PET back to BTX by unlocking
in plastic waste exceeds efforts to mitigate Lucena, R., Cardenas, S., Sans, V., and de hidden hydrogen. ChemSusChem. Published
plastic pollution. Science 369, 1515–1518. Pedro, I. (2020). Paramagnetic ionic liquid- online March 4, 2021. https://doi.org/10.
coated SiO2@Fe3O4 nanoparticles—The 1002/cssc.202100196.
54. Esposito, D. (2019). Plastic upcycling. Nat. next generation of magnetically recoverable
Catal. 2, 945–946. nanocatalysts applied in the glycolysis of PET. 81. Vollmer, I., Jenks, M.J.F., González, R.M.,
Appl. Catal. B 260, 118110. Meirer, F., and Weckhuysen, B.M. (2021).
55. Sohn, Y.J., Kim, H.T., Baritugo, K.-A., Jo, S.Y., Plastic waste conversion over a refinery waste
Song, H.M., Park, S.Y., Park, S.K., Pyo, J., Cha, 69. Sheel, A., and Pant, D. (2018). Chemical catalyst. Angew. Chem. Int. Ed. Engl.
H.G., Kim, H., et al. (2020). Recent advances in depolymerization of polyurethane foams via Published online May 11, 2021. https://doi.
sustainable plastic upcycling and glycolysis and hydrolysis. In Recycling of org/10.1002/anie.202104110.
biopolymers. Biotechnol. J. 15, e1900489. Polyurethane Foams, S. Thomas, A.V. Rane, K.
Kanny, A. Vk, and M.G. Thomas, eds. (William 82. Bäckström, E., Odelius, K., and Hakkarainen,
56. Hou, Q., Qi, X., Zhen, M., Qian, H., Nie, Y., Bai,
Andrew Publishing), pp. 67–75. M. (2019). Designed from recycled: turning
C., Zhang, S., Bai, X., and Ju, M. (2021). polyethylene waste to covalently attached
Biorefinery roadmap based on catalytic 70. Krall, E.M., Klein, T.W., Andersen, R.J., Nett, polylactide plasticizers. ACS Sustain. Chem.
production and upgrading of 5- A.J., Glasgow, R.W., Reader, D.S., Eng. 7, 11004–11013.
hydroxymethylfurfural. Green Chem. 23, Dauphinais, B.C., Mc Ilrath, S.P., Fischer, A.A.,
119–231. Carney, M.J., et al. (2014). Controlled 83. Bäckström, E., Odelius, K., and Hakkarainen,
hydrogenative depolymerization of M. (2017). Trash to treasure: microwave-
57. Weckhuysen, B.M. (2020). Creating value from
polyesters and polycarbonates catalyzed by assisted conversion of polyethylene to
plastic waste. Science 370, 400–401.
ruthenium(II) PNN pincer complexes. Chem. functional chemicals. Ind. Eng. Chem. Res. 56,
58. Fendall, L.S., and Sewell, M.A. (2009). Commun. (Camb.) 50, 4884–4887. 14814–14821.
Contributing to marine pollution by washing
71. Fuentes, J.A., Smith, S.M., Scharbert, M.T., 84. Beydoun, K., and Klankermayer, J. (2020).
your face: microplastics in facial cleansers.
Carpenter, I., Cordes, D.B., Slawin, A.M.Z., Efficient plastic waste recycling to value-
Mar. Pollut. Bull. 58, 1225–1228.
and Clarke, M.L. (2015). On the functional added products by integrated biomass
59. Sun, J., Dai, X., Wang, Q., van Loosdrecht, group tolerance of ester hydrogenation and processing. ChemSusChem 13, 488–492.
M.C.M., and Ni, B.-J. (2019). Microplastics in polyester depolymerisation catalysed by
ruthenium complexes of tridentate 85. Wong, S.L., Ngadi, N., Abdullah, T.A.T., and
wastewater treatment plants: Detection,
occurrence and removal. Water Res. 152, aminophosphine ligands. Chemistry 21, Inuwa, I.M. (2015). Current state and future
10851–10860. prospects of plastic waste as source of fuel: a
21–37.
review. Renew. Sustain. Energy Rev. 50, 1167–
60. Hou, L., Kumar, D., Yoo, C.G., Gitsov, I., and 72. Westhues, S., Idel, J., and Klankermayer, J. 1180.
Majumder, E.L.W. (2021). Conversion and (2018). Molecular catalyst systems as key
removal strategies for microplastics in enablers for tailored polyesters and 86. Uekert, T., Pichler, C.M., Schubert, T., and
wastewater treatment plants and landfills. polycarbonate recycling concepts. Sci. Adv. 4, Reisner, E. (2020). Solar-driven reforming of
Chem. Eng. J. 406, 126715. eaat9669. solid waste for a sustainable future. Nat.
Sustain. Published online November 30, 2020.
61. Min, K., Cuiffi, J.D., and Mathers, R.T. (2020). 73. Jing, Y., Wang, Y., Furukawa, S., Xia, J., Sun, https://doi.org/10.1038/s41893-020-00650-x.
Ranking environmental degradation trends of C., Hülsey, M.J., Wang, H., Guo, Y., Liu, X.,
plastic marine debris based on physical and Yan, N. (2021). Towards the circular 87. Chin, K.F., Ðokic, M., and Soo, H.S. (2020).
properties and molecular structure. Nat. economy: converting aromatic plastic waste Artificial photosynthesis beyond water
Commun. 11, 727. back to arenes over a Ru/Nb2 O5 catalyst. splitting for environmental remediation.
Angew. Chem. Int. Ed. Engl. 60, 5527–5535. Trends Chem. 2, 485–488.
62. Ru, J., Huo, Y., and Yang, Y. (2020). Microbial
degradation and valorization of plastic 74. Kumar, A., von Wolff, N., Rauch, M., Zou, 88. Tomoji, K., and Tadayoshi, S. (1981).
wastes. Front. Microbiol. 11, 442. Y.-Q., Shmul, G., Ben-David, Y., Leitus, G., Photocatalytic hydrogen production from
Avram, L., and Milstein, D. (2020). water by the decomposition of poly-
63. Shabbir, S., Faheem, M., Ali, N., Kerr, P.G., Hydrogenative depolymerization of nylons. vinylchloride, protein, algae, dead insects,
Wang, L.-F., Kuppusamy, S., and Li, Y. (2020). J. Am. Chem. Soc. 142, 14267–14275. and excrement. Chem. Lett. 10, 81–84.
Periphytic biofilm: an innovative approach for
biodegradation of microplastics. Sci. Total 75. Woo, O.S., Kruse, T.M., and Broadbelt, L.J. 89. Uekert, T., Kuehnel, M.F., Wakerley, D.W.,
Environ. 717, 137064. (2000). Binary mixture pyrolysis of polystyrene and Reisner, E. (2018). Plastic waste as a
and poly(a-methylstyrene). Polym. Degrad. feedstock for solar-driven H2 generation.
64. Kang, J., Zhou, L., Duan, X., Sun, H., Ao, Z., Stabil. 70, 155–160. Energy Environ. Sci. 11, 2853–2857.
and Wang, S. (2019). Degradation of cosmetic
microplastics via functionalized carbon 76. Wang, Y., Hou, Y., and Song, H. (2017). Ring- 90. Wakerley, D.W., Kuehnel, M.F., Orchard, K.L.,
nanosprings. Matter 1, 745–758. closing depolymerization of Ly, K.H., Rosser, T.E., and Reisner, E. (2017).
polytetrahydrofuran to produce Solar-driven reforming of lignocellulose to H2
65. Jiao, X., Zheng, K., Chen, Q., Li, X., Li, Y., Shao, tetrahydrofuran using heteropolyacid as with a CdS/CdOx photocatalyst. Nat. Energy
W., Xu, J., Zhu, J., Pan, Y., Sun, Y., and Xie, Y. catalyst. Polym. Degrad. Stabil. 144, 17–23. 2, 17021.

Cell Reports Physical Science 2, 100514, August 18, 2021 27


ll
OPEN ACCESS
Review

91. Uekert, T., Kasap, H., and Reisner, E. (2019). interaction in A-site-deficient perovskite via Ni 116. Liu, X., Ma, C., Wen, Y., Chen, X., Zhao, X.,
Photoreforming of nonrecyclable plastic substitution to promote the thermocatalytic Tang, T., Holze, R., and Mijowska, E. (2021).
waste over a carbon nitride/nickel phosphide synthesis of carbon nanotubes from plastics. Highly efficient conversion of waste plastic
catalyst. J. Am. Chem. Soc. 141, 15201–15210. J. Hazard. Mater. 403, 123642. into thin carbon nanosheets for superior
capacitive energy storage. Carbon 171,
92. Pichler, C.M., Uekert, T., and Reisner, E. 104. Liu, X., He, S., Han, Z., and Wu, C. (2021). 819–828.
(2020). Photoreforming of biomass in metal Investigation of spherical alumina supported
salt hydrate solutions. Chem. Commun. catalyst for carbon nanotubes production 117. Kim, P.J., Fontecha, H.D., Kim, K., and Pol,
(Camb.) 56, 5743–5746. from waste polyethylene. Process Saf. V.G. (2018). Toward high-performance
Environ. Prot. 146, 201–207. lithium-sulfur batteries: upcycling of LDPE
93. Dufaud, V., and Basset, J.-M. (1998). Catalytic plastic into sulfonated carbon scaffold via
hydrogenolysis at low temperature and 105. Yang, R.-X., Wu, S.-L., Chuang, K.-H., and microwave-promoted sulfonation. ACS Appl.
pressure of polyethylene and polypropylene Wey, M.-Y. (2020). Co-production of carbon Mater. Interfaces 10, 14827–14834.
to diesels or lower alkanes by a zirconium nanotubes and hydrogen from waste plastic
hydride supported on silica-alumina: a step gasification in a two-stage fluidized catalytic 118. Kang, M.J., Kim, H.T., Lee, M.-W., Kim, K.-A.,
toward polyolefin degradation by the bed. Renew. Energy 159, 10–22. Khang, T.U., Song, H.M., Park, S.J., Joo, J.C.,
microscopic reverse of Ziegler-Natta and Cha, H.G. (2020). A chemo-microbial
polymerization. Angew. Chem. Int. Ed. Engl. 106. Ko, S., Kwon, Y.J., Lee, J.U., and Jeon, Y.-P. hybrid process for the production of 2-
37, 806–810. (2020). Preparation of synthetic graphite from pyrone-4,6-dicarboxylic acid as a promising
waste PET plastic. J. Ind. Eng. Chem. 83, bioplastic monomer from PET waste. Green
94. Celik, G., Kennedy, R.M., Hackler, R.A., 449–458. Chem. 22, 3461–3469.
Ferrandon, M., Tennakoon, A., Patnaik, S.,
LaPointe, A.M., Ammal, S.C., Heyden, A., 107. Pandey, S., Karakoti, M., Dhali, S., Karki, N., 119. Jones, G.O., Yuen, A., Wojtecki, R.J., Hedrick,
Perras, F.A., et al. (2019). Upcycling single-use SanthiBhushan, B., Tewari, C., Rana, S., J.L., and Garcı́a, J.M. (2016). Computational
polyethylene into high-quality liquid Srivastava, A., Melkani, A.B., and Sahoo, N.G. and experimental investigations of one-step
products. ACS Cent. Sci. 5, 1795–1803. (2019). Bulk synthesis of graphene nanosheets conversion of poly(carbonate)s into value-
from plastic waste: An invincible method of added poly(aryl ether sulfone)s. Proc. Natl.
95. Liu, S., Kots, P.A., Vance, B.C., Danielson, A., solid waste management for better Acad. Sci. USA 113, 7722–7726.
and Vlachos, D.G. (2021). Plastic waste to fuels tomorrow. Waste Manag. 88, 48–55.
by hydrocracking at mild conditions. Sci. Adv. 120. Gardea, F., Garcia, J.M., Boday, D.J., Bajjuri,
7, eabf8283. 108. Luong, D.X., Bets, K.V., Algozeeb, W.A., K.M., Naraghi, M., and Hedrick, J.L. (2014).
Stanford, M.G., Kittrell, C., Chen, W., Hybrid poly(aryl ether sulfone amide)s for
96. Jia, C., Xie, S., Zhang, W., Intan, N.N., Salvatierra, R.V., Ren, M., McHugh, E.A., advanced thermoplastic composites.
Sampath, J., Pfaendtner, J., and Lin, H. (2021). Advincula, P.A., et al. (2020). Gram-scale Macromol. Chem. Phys. 215, 2260–2267.
Deconstruction of high-density polyethylene bottom-up flash graphene synthesis. Nature
into liquid hydrocarbon fuels and lubricants 577, 647–651. 121. Demarteau, J., O’Harra, K.E., Bara, J.E., and
by hydrogenolysis over Ru catalyst. Chem Sardon, H. (2020). Valorization of plastic
Catal. Published online May 17, 2021. https:// 109. Bazargan, A., and McKay, G. (2012). A wastes for the synthesis of imidazolium-based
doi.org/10.1016/j.checat.2021.04.002. review—synthesis of carbon nanotubes from self-supported elastomeric ionenes.
plastic wastes. Chem. Eng. J. 195-196, ChemSusChem 13, 3122–3126.
97. Chen, S., Liu, Z., Jiang, S., and Hou, H. (2020).
377–391.
Carbonization: a feasible route for 122. Quaranta, E., Dibenedetto, A., Nocito, F., and
reutilization of plastic wastes. Sci. Total 110. Yao, D., Zhang, Y., Williams, P.T., Yang, H., Fini, P. (2021). Chemical recycling of poly-
Environ. 710, 136250. and Chen, H. (2018). Co-production of (bisphenol A carbonate) by diaminolysis: a
98. Chaudhary, S., Kumari, M., Chauhan, P., and hydrogen and carbon nanotubes from real- new carbon-saving synthetic entry into non-
Ram Chaudhary, G. (2021). Upcycling of world waste plastics: Influence of catalyst isocyanate polyureas (NIPUreas). J. Hazard.
plastic waste into fluorescent carbon dots: an composition and operational parameters. Mater. 403, 123957.
environmentally viable transformation to Appl. Catal. B 221, 584–597.
123. Saito, K., Jehanno, C., Meabe, L., Olmedo-
biocompatible C-dots with potential
111. Algozeeb, W.A., Savas, P.E., Luong, D.X., Martı́nez, J.L., Mecerreyes, D., Fukushima, K.,
prospective in analytical applications. Waste Chen, W., Kittrell, C., Bhat, M., Shahsavari, R., and Sardon, H. (2020). From plastic waste to
Manag. 120, 675–686.
and Tour, J.M. (2020). Flash graphene from polymer electrolytes for batteries through
99. Gong, J., Liu, J., Jiang, Z., Chen, X., Wen, X., plastic waste. ACS Nano 14, 15595–15604. chemical upcycling of polycarbonate.
Mijowska, E., and Tang, T. (2014). Converting J. Mater. Chem. A Mater. Energy Sustain. 8,
112. Yuan, X., Cho, M.-K., Lee, J.G., Choi, S.W., and 13921–13926.
mixed plastics into mesoporous hollow
Lee, K.B. (2020). Upcycling of waste
carbon spheres with controllable diameter.
polyethylene terephthalate plastic bottles 124. Demarteau, J., Olazabal, I., Jehanno, C., and
Appl. Catal. B 152-153, 289–299.
into porous carbon for CF4 adsorption. Sardon, H. (2020). Aminolytic upcycling of
100. Wei, T., Zhang, Z., Zhu, Z., Zhou, X., Wang, Y., Environ. Pollut. 265 (pt A), 114868. poly(ethylene terephthalate) wastes using a
Wang, Y., and Zhuang, Q. (2019). Recycling of thermally-stable organocatalyst. Polym.
waste plastics and scalable preparation of Si/ 113. Daniel, G., Kosmala, T., Dalconi, M.C., Nodari, Chem. 11, 4875–4882.
CNF/C composite as anode material for L., Badocco, D., Pastore, P., Lorenzetti, A.,
lithium-ion batteries. Ionics 25, 1523–1529. Granozzi, G., and Durante, C. (2020). 125. Fukushima, K., Jones, G.O., Horn, H.W., Rice,
Upcycling of polyurethane into iron-nitrogen- J.E., Kato, T., and Hedrick, J.L. (2020).
101. Choi, D., Yeo, J.-S., Joh, H.-I., and Lee, S. carbon electrocatalysts active for oxygen Formation of bis-benzimidazole and bis-
(2018). Carbon nanosheet from polyethylene reduction reaction. Electrochim. Acta 362, benzoxazole through organocatalytic
thin film as a transparent conducting film: 137200. depolymerization of poly(ethylene
‘‘upcycling’’ of waste to organic photovoltaics terephthalate) and its mechanism. Polym.
application. ACS Sustain. Chem. Eng. 6, 114. Wen, Y., Kierzek, K., Chen, X., Gong, J., Liu, J., Chem. 11, 4904–4913.
12463–12470. Niu, R., Mijowska, E., and Tang, T. (2019).
Mass production of hierarchically porous 126. Scheutz, G.M., Lessard, J.J., Sims, M.B., and
102. Min, J., Wen, X., Tang, T., Chen, X., Huo, K., carbon nanosheets by carbonizing ‘‘real- Sumerlin, B.S. (2019). Adaptable crosslinks in
Gong, J., Azadmanjiri, J., He, C., and world’’ mixed waste plastics toward excellent- polymeric materials: resolving the intersection
Mijowska, E. (2020). A general approach performance supercapacitors. Waste Manag. of thermoplastics and thermosets. J. Am.
towards carbonization of plastic waste into a 87, 691–700. Chem. Soc. 141, 16181–16196.
well-designed 3D porous carbon framework
for super lithium-ion batteries. Chem. 115. Liu, N., Hao, L., Zhang, B., Niu, R., Gong, J., 127. Podgórski, M., Fairbanks, B.D., Kirkpatrick,
Commun. (Camb.) 56, 9142–9145. and Tang, T. (2020). High-performance solar B.E., McBride, M., Martinez, A., Dobson, A.,
vapor generation by sustainable biomimetic Bongiardina, N.J., and Bowman, C.N. (2020).
103. Jia, J., Veksha, A., Lim, T.-T., and Lisak, G. snake-scale-like porous carbon. Sustain. Covalent adaptable networks: toward stimuli-
(2021). Weakening the strong Fe-La Energy Fuels 4, 5522–5532. responsive dynamic thermosets through

28 Cell Reports Physical Science 2, 100514, August 18, 2021


ll
OPEN ACCESS
Review

continuous development and improvements modification using radiation-induced 153. Lessard, J.J., Scheutz, G.M., Sung, S.H., Lantz,
in covalent adaptable networks (CANs). Adv. grafting. Appl. Surf. Sci. 422, 720–730. K.A., Epps, T.H., and Sumerlin, B.S. (2020).
Mater. 32, 2070158. Block copolymer vitrimers. J. Am. Chem. Soc.
141. Liu, X., Hong, M., Falivene, L., Cavallo, L., and 283–289.
128. Montarnal, D., Capelot, M., Tournilhac, F., Chen, E.Y.X. (2019). Closed-loop polymer
and Leibler, L. (2011). Silica-like malleable upcycling by installing property-enhancing 154. Chen, Y., Wu, X., Wei, J., Wu, H., and Fang, J.
materials from permanent organic networks. comonomer sequences and recyclability. (2020). Reuse polyester/cotton blend fabrics
Science 334, 965–968. Macromolecules 52, 4570–4578. to prepare fiber reinforced composite:
fabrication, characterization, and interfacial
129. Elling, B.R., and Dichtel, W.R. (2020). 142. Eagan, J.M., Xu, J., Di Girolamo, R., Thurber, properties evaluation. Polym. Compos.
Reprocessable cross-linked polymer C.M., Macosko, C.W., LaPointe, A.M., Bates, Published online September 9, 2020. https://
networks: are associative exchange F.S., and Coates, G.W. (2017). Combining doi.org/10.1002/pc.25814.
mechanisms desirable? ACS Cent. Sci. 6, polyethylene and polypropylene: enhanced
1488–1496. performance with PE/iPP multiblock 155. Sharma, K., Khilari, V., Chaudhary, B.U., Jogi,
polymers. Science 355, 814–816. A.B., Pandit, A.B., and Kale, R.D. (2020).
130. Röttger, M., Domenech, T., van der Weegen, Cotton based composite fabric reinforced
R., Breuillac, A., Nicolaÿ, R., and Leibler, L. 143. Cruz Sanchez, F.A., Boudaoud, H., Camargo, with waste polyester fibers for improved
(2017). High-performance vitrimers from M., and Pearce, J.M. (2020). Plastic recycling in mechanical properties. Waste Manag. 107,
commodity thermoplastics through additive manufacturing: a systematic 227–234.
dioxaborolane metathesis. Science 356, literature review and opportunities for the
62–65. circular economy. J. Clean. Prod. 264, 121602. 156. Dardeer, H.M., and Toghan, A. (2021). A novel
route for the synthesis of pseudopolyrotaxane
131. Kar, G.P., Saed, M.O., and Terentjev, E.M. 144. Mejia, E.B., Al-Maqdi, S., Alkaabi, M., containing g-Cyclodextrin based on
(2020). Scalable upcycling of thermoplastic Alhammadi, A., Alkaabi, M., Cherupurakal, N., environmental waste recycling. J. Mol. Struct.
polyolefins into vitrimers through and Mourad, A.I. (2020). Upcycling of HDPE 1227, 129707.
transesterification. J. Mater. Chem. A Mater. waste using additive manufacturing:
Energy Sustain. 8, 24137–24147. feasibility and challenges. In Advances in 157. El-Sayed, E.-S.M., and Yuan, D. (2020). Waste
Science and Engineering Technology to MOFs: sustainable linker, metal, and
132. Sheppard, D.T., Jin, K., Hamachi, L.S., Dean, solvent sources for value-added MOF
International Conferences (Curran
W., Fortman, D.J., Ellison, C.J., and Dichtel, synthesis and applications. Green Chem. 22,
Associates), pp. 1–6.
W.R. (2020). Reprocessing postconsumer 4082–4104.
polyurethane foam using carbamate 145. Nur-A-Tomal, M.S., Pahlevani, F., and
exchange catalysis and twin-screw extrusion. 158. Slater, B., Wong, S.-O., Duckworth, A., White,
Sahajwalla, V. (2020). Direct transformation of
ACS Cent. Sci. 6, 921–927. A.J.P., Hill, M.R., and Ladewig, B.P. (2019).
waste children’s toys to high quality products
Upcycling a plastic cup: one-pot synthesis of
using 3D printing: a waste-to-wealth and
133. Breuillac, A., Caffy, F., Vialon, T., and Nicolaÿ, lactate containing metal organic frameworks
sustainable approach. J. Clean. Prod. 267,
R. (2020). Functionalization of polyisoprene from polylactic acid. Chem. Commun.
122188.
and polystyrene via reactive processing using (Camb.) 55, 7319–7322.
azidoformate grafting agents, and its 146. Kamleitner, F., Duscher, B., Koch, T., Knaus,
application to the synthesis of dioxaborolane- 159. Semyonov, O., Chaemchuen, S., Ivanov, A.,
S., and Archodoulaki, V.M. (2017). Upcycling Verpoort, F., Kolska, Z., Syrtanov, M., Svorcik,
based polyisoprene vitrimers. Polym. Chem. of polypropylene—the influence of
11, 6479–6491. V., Yusubov, M.S., Lyutakov, O., Guselnikova,
polyethylene impurities. Polym. Eng. Sci. 57, O., and Postnikov, P.S. (2021). Smart recycling
134. Lessard, J.J., Garcia, L.F., Easterling, C.P., 1374–1381. of PET to sorbents for insecticides through
Sims, M.B., Bentz, K.C., Arencibia, S., Savin, in situ MOF growth. Appl. Mater. Today 22,
147. Creton, C. (2017). Molecular stitches for
D.A., and Sumerlin, B.S. (2019). Catalyst-free 100910.
enhanced recycling of packaging. Science
vitrimers from vinyl polymers.
355, 797–798. 160. Ren, W., Zhou, P., Nie, G., Cheng, C., Duan,
Macromolecules 52, 2105–2111.
X., Zhang, H., and Wang, S. (2020). Hydroxyl
148. Luzuriaga, S.E., Kovárová, J., and Fortelný, I. radical dominated elimination of plasticizers
135. Williamson, J.B., Lewis, S.E., Johnson, R.R.,
(2011). Stability of model recycled mixed by peroxymonosulfate on metal-free boron:
3rd, Manning, I.M., and Leibfarth, F.A. (2019).
plastic waste compatibilised with a Kinetics and mechanisms. Water Res. 186,
C-H Functionalization of commodity
cooperative compatibilisation system. Polym. 116361.
polymers. Angew. Chem. Int. Ed. Engl. 58,
Degrad. Stabil. 96, 751–755.
8654–8668.
161. Wang, B., Wang, X., Xu, N., Shen, Y., Lu, F.,
136. Lewis, S.E., Wilhelmy, B.E., Jr., and Leibfarth, 149. Li, H., Sui, X., and Xie, X.-M. (2017). High- Liu, Y., Huang, Y., and Hu, Z. (2020). Recycling
F.A. (2019). Upcycling aromatic polymers strength and super-tough PA6/PS/PP/SEBS of carbon fibers from unsaturated polyester
through C–H fluoroalkylation. Chem. Sci. quaternary blends compatibilized by using a composites via a hydrolysis-oxidation
(Camb.) 10, 6270–6277. highly effective multi-phase compatibilizer: synergistic catalytic strategy. Compos. Sci.
toward efficient recycling of waste plastics. Technol. d203, 108589.
137. Easterling, C.P., Kubo, T., Orr, Z.M., Fanucci, Polymer (Guildf.) 123, 240–246.
G.E., and Sumerlin, B.S. (2017). Synthetic 162. Miao, F., Liu, Y., Gao, M., Yu, X., Xiao, P.,
upcycling of polyacrylates through 150. Jiang, H., Ye, L., Wang, Y., Ma, L., Cui, D., and Wang, M., Wang, S., and Wang, X. (2020).
organocatalyzed post-polymerization Tang, T. (2020). Synthesis and characterization Degradation of polyvinyl chloride
modification. Chem. Sci. (Camb.) 8, 7705– of polypropylene-based polyurethanes. microplastics via an electro-Fenton-like
7709. Macromolecules 53, 3349–3357. system with a TiO2/graphite cathode.
J. Hazard. Mater. 399, 123023.
138. Van Guyse, J.F.R., Verjans, J., Vandewalle, S., 151. Nomura, K., Peng, X., Kim, H., Jin, K., Kim,
De Bruycker, K., Du Prez, F.E., and H.J., Bratton, A.F., Bond, C.R., Broman, A.E., 163. Tofa, T.S., Ye, F., Kunjali, K.L., and Dutta, J.
Hoogenboom, R. (2019). Full and partial Miller, K.M., and Ellison, C.J. (2020). (2019). Enhanced visible light
amidation of poly(methyl acrylate) as basis for Multiblock copolymers for recycling photodegradation of microplastic fragments
functional polyacrylamide (co)polymers. polyethylene-poly(ethylene terephthalate) with plasmonic platinum/zinc oxide nanorod
Macromolecules 52, 5102–5109. mixed waste. ACS Appl. Mater. Interfaces 12, photocatalysts. Catalysts 9, 819.
9726–9735.
139. Jeon, C., Kim, D.W., Chang, S., Kim, J.G., and 164. Llorente-Garcı́a, B.E., Hernández-López, J.M.,
Seo, M. (2019). Synthesis of polypropylene via 152. Matias, Á.A., Lima, M.S., Pereira, J., Pereira, Zaldı́var-Cadena, A.A., Siligardi, C., and
catalytic deoxygenation of poly(methyl P., Barros, R., Coelho, J.F.J., and Serra, A.C. Cedillo-González, E.I. (2020). First insights
acrylate). ACS Macro Lett. 8, 1172–1178. (2020). Use of recycled polypropylene/ into photocatalytic degradation of HDPE and
poly(ethylene terephthalate) blends to LDPE microplastics by a mesoporous N–TiO2
140. Hassan, M., Taimur, S., and Yasin, T. (2017). manufacture water pipes: an industrial scale coating: effect of size and shape of
Upcycling of polypropylene waste by surface study. Waste Manag. 101, 250–258. microplastics. Coatings 10, 658.

Cell Reports Physical Science 2, 100514, August 18, 2021 29


ll
OPEN ACCESS
Review

165. Jiang, R., Lu, G., Yan, Z., Liu, J., Wu, D., and 173. Garcı́a, J.M., Jones, G.O., Virwani, K., 181. Debsharma, T., Yagci, Y., and Schlaad, H.
Wang, Y. (2021). Microplastic degradation by McCloskey, B.D., Boday, D.J., ter Huurne, (2019). Cellulose-derived functional
hydroxy-rich bismuth oxychloride. J. Hazard. G.M., Horn, H.W., Coady, D.J., Bintaleb, A.M., polyacetal by cationic ring-opening
Mater. 405, 124247. Alabdulrahman, A.M.S., et al. (2014). polymerization of levoglucosenyl methyl
Recyclable, strong thermosets and ether. Angew. Chem. Int. Ed. Engl. 58, 18492–
166. Gazi, S., Ðokic, M., Chin, K.F., Ng, P.R., and organogels via paraformaldehyde 18495.
Soo, H.S. (2019). Visible light-driven cascade condensation with diamines. Science 344,
carbon-carbon bond scission for organic 732–735. 182. Tang, X., Westlie, A.H., Caporaso, L., Cavallo,
transformations and plastics recycling. Adv. L., Falivene, L., and Chen, E.Y.X. (2020).
Sci. (Weinh.) 6, 1902020. 174. Shieh, P., Zhang, W., Husted, K.E.L., Kristufek, Biodegradable polyhydroxyalkanoates by
S.L., Xiong, B., Lundberg, D.J., Lem, J., stereoselective copolymerization of racemic
167. Walker, T.W., Frelka, N., Shen, Z., Chew, A.K., Veysset, D., Sun, Y., Nelson, K.A., et al. (2020). diolides: stereocontrol and polyolefin-like
Banick, J., Grey, S., Kim, M.S., Dumesic, J.A., Cleavable comonomers enable degradable, properties. Angew. Chem. Int. Ed. Engl. 59,
Van Lehn, R.C., and Huber, G.W. (2020). recyclable thermoset plastics. Nature 583, 7881–7890.
Recycling of multilayer plastic packaging 542–547.
materials by solvent-targeted recovery and 183. Maiti, S., Manna, S., Banahene, N., Pham, L.,
precipitation. Sci. Adv. 6, eaba7599. 175. Jin, Y., Lei, Z., Taynton, P., Huang, S., and Liang, Z., Wang, J., Xu, Y., Bettinger, R.,
Zhang, W. (2019). Malleable and recyclable Zientko, J., Esser-Kahn, A.P., and Du, W.
168. Jing, Y., Wang, Y., Furukawa, S., Xia, J., Sun, thermosets: the next generation of plastics. (2020). From glucose to polymers: a
C., Hülsey, M.J., Wang, H., Guo, Y., Liu, X., and Matter 1, 1456–1493. continuous chemoenzymatic process. Angew.
Yan, N. (2021). Towards the circular economy: Chem. Int. Ed. Engl. 59, 18943–18947.
converting aromatic plastic waste back to 176. Wang, D., Cui, J., Gan, M., Xue, Z., Wang, J.,
arenes over Ru/Nb2O5 catalyst. Angew. Liu, P., Hu, Y., Pardo, Y., Hamada, S., Yang, D., 184. Bhaumik, A., Peterson, G.I., Kang, C., and
Chem. Int. Ed. Engl. Published online March 1, et al. (2020). Transformation of biomass dna Choi, T.-L. (2019). Controlled living cascade
2021. https://doi.org/10.1002/anie. into biodegradable materials from gels to polymerization to make fully degradable
202011063. plastics for reducing petrochemical sugar-based polymers from D-glucose and D-
consumption. J. Am. Chem. Soc. 142, 10114– galactose. J. Am. Chem. Soc. 141, 12207–
10124. 12211.
169. Dong, L., Lin, L., Han, X., Si, X., Liu, X., Guo, Y.,
Lu, F., Rudic, S., Parker, S.F., Yang, S., and 185. Zhang, Q., Deng, Y., Shi, C.-Y., Feringa, B.L.,
Wang, Y. (2019). Breaking the limit of lignin 177. Ates, B., Koytepe, S., Ulu, A., Gurses, C., and Tian, H., and Qu, D.-H. (2021). Dual closed-
monomer production via cleavage of interunit Thakur, V.K. (2020). Chemistry, structures, and loop chemical recycling of synthetic polymers
carbon–carbon linkages. Chem 5, 1521–1536. advanced applications of nanocomposites by intrinsically reconfigurable poly(disulfides).
from biorenewable resources. Chem. Rev. Matter 4, 1352–1364.
170. Zhou, Y., Shan, Y., Guan, D., Liang, X., Cai, Y., 120, 9304–9362.
Liu, J., Xie, W., Xue, J., Ma, Z., and Yang, Z. 186. Häußler, M., Eck, M., Rothauer, D., and
(2020). Sharing tableware reduces waste 178. Tang, X., Westlie, A.H., Watson, E.M., and Mecking, S. (2021). Closed-loop recycling of
generation, emissions and water Chen, E.Y.-X. (2019). Stereosequenced polyethylene-like materials. Nature 590,
consumption in China’s takeaway packaging crystalline polyhydroxyalkanoates from 423–427.
waste dilemma. Nat. Food 1, 552–561. diastereomeric monomer mixtures. Science
366, 754–758. 187. Yuan, J., Xiong, W., Zhou, X., Zhang, Y., Shi,
171. Zhu, J.-B., Watson, E.M., Tang, J., and Chen, D., Li, Z., and Lu, H. (2019). 4-hydroxyproline-
E.Y.-X. (2018). A synthetic polymer system with 179. Liu, Y., and Mecking, S. (2019). A synthetic derived sustainable polythioesters: controlled
repeatable chemical recyclability. Science polyester from plant oil feedstock by ring-opening polymerization, complete
360, 398–403. functionalizing polymerization. Angew. recyclability, and facile functionalization.
Chem. Int. Ed. Engl. 58, 3346–3350. J. Am. Chem. Soc. 141, 4928–4935.
172. Mohadjer Beromi, M., Kennedy, C.R.,
Younker, J.M., Carpenter, A.E., Mattler, S.J., 180. Debsharma, T., Behrendt, F.N., Laschewsky, 188. Rizzo, A., Peterson, G.I., Bhaumik, A., Kang,
Throckmorton, J.A., and Chirik, P.J. (2021). A., and Schlaad, H. (2019). Ring-opening C., and Choi, T.L. (2020). Sugar-based
Iron-catalysed synthesis and chemical metathesis polymerization of biomass- polymers from d-xylose: living cascade
recycling of telechelic 1,3-enchained derived levoglucosenol. Angew. Chem. Int. polymerization. Angew. Chem. Int. Ed. Engl.
oligocyclobutanes. Nat. Chem. 13, 156–162. Ed. Engl. 58, 6718–6721. 60, 849–855.

30 Cell Reports Physical Science 2, 100514, August 18, 2021

You might also like