Sei sulla pagina 1di 47

Review

pubs.acs.org/CR

Catalysis as an Enabling Science for Sustainable Polymers


Xiangyi Zhang,† Mareva Fevre,‡ Gavin O. Jones,‡ and Robert M. Waymouth*,†

Department of Chemistry, Stanford University, Stanford, California 94305-5080, United States

IBM Research−Almaden, 650 Harry Road, San Jose, California 95120, United States

ABSTRACT: The replacement of current petroleum-based plastics with sustainable


alternatives is a crucial but formidable challenge for the modern society. Catalysis
presents an enabling tool to facilitate the development of sustainable polymers. This
review provides a system-level analysis of sustainable polymers and outlines key criteria
with respect to the feedstocks the polymers are derived from, the manner in which the
polymers are generated, and the end-of-use options. Specifically, we define sustainable
polymers as a class of materials that are derived from renewable feedstocks and exhibit
closed-loop life cycles. Among potential candidates, aliphatic polyesters and
polycarbonates are promising materials due to their renewable resources and excellent
biodegradability. The development of renewable monomers, the versatile synthetic routes
to convert these monomers to polyesters and polycarbonate, and the different end-of-use
options for these polymers are critically reviewed, with a focus on recent advances in
catalytic transformations that lower the technological barriers for developing more sustainable replacements for petroleum-based
plastics.

CONTENTS 4.3. Biological Recycling 868


4.3.1. Anaerobic Bacteria 869
1. Introduction 839 4.3.2. Aerobic Bacteria 869
1.1. Definition 840 4.3.3. Cells Extracts 870
1.2. Scope of Review 840 5. Conclusions and Outlook 871
1.3. General Principles 841 Author Information 872
1.3.1. Feedstocks, Monomers, and Polymers 841 Corresponding Author 872
1.3.2. End-of-Use Options 841 ORCID 872
1.4. Challenges 841 Notes 872
1.5. Metrics Used to Evaluate Sustainability 842 Biographies 872
1.5.1. Green Design Metrics (GDM) 842 Acknowledgments 872
1.5.2. Life Cycle Assessment (LCA) 842 References 872
1.5.3. Relation Between the Two Metrics 843 Note Added after ASAP Publication 885
1.6. Commercial Examples 843
1.7. Opportunities and Challenges for the Devel-
opment of Sustainable Polyesters and
Polycarbonates 843 1. INTRODUCTION
2. Renewable Monomers 844 The development of petroleum-based plastics is one of the
2.1. Diacids, Hydroxyl Acids 845 crowning achievements of the 20th Century. Eighty years of
2.2. Polyols 847 commercial development have led to a family of materials
2.3. Cyclic Esters 847 whose low cost, processing versatility, and range of mechanical
2.4. Carbonates 848 properties have made them the materials of everyday life, from
2.5. Epoxides 850 clothing, cutlery, automobiles to healthcare, electronics, and the
3. Synthesis of Polyesters and Polycarbonates 850 materials of our modern infrastructure. From 1.65 million tons
3.1. Step-Growth Polymerization (SGP) 851 in 1950 to 311 million tons in 2014 worldwide,1 plastics usage
3.2. Chain-Growth Polymerization 853 is expanding and expected to grow at a steady pace of 3−4%
3.2.1. Ring-Opening Polymerization (ROP) 853 per year.2 The relatively low costs of energy, abundance of
3.2.2. Ring-Opening Copolymerization (ROCP) 859 waste disposal sites, and an economic and cultural focus on the
3.3. Microbial Fermentation 861 convenience of disposable items of commerce have fostered an
4. End-of-Use Options 862
4.1. Mechanical Recycling 863
Special Issue: Sustainable Chemistry
4.2. Chemical Recycling 863
4.2.1. Solvolysis 863 Received: June 8, 2017
4.2.2. Thermal Recycling (Pyrolysis) 866 Published: October 19, 2017

© 2017 American Chemical Society 839 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

industrial ecology based on a linear model of resource science, an effort to provide a coherent system-level analysis to
utilization (Resource-Monomer-Polymer-Product-Waste, Fig- define sustainability and to identify the appropriate goals,
ure 1, black lines).3 Petrochemical feedstocks that provide the indices, and metrics to assess whether a particular activity is
sustainable.7−10 In this context, the definition of sustainable
polymers should be considered at a system level, encompassing
considerations such as the resources from which they are made,
the manner in which they are produced, stored, distributed,
utilized, and either recovered or otherwise released into the
environment. Concepts from industrial ecology3 and circular
economy11 can provide a useful framework for the definition of
a sustainable polymer, that is a class of materials that exhibit
closed-loop life cycles, such that they are components of a
system where “resources” and “waste” are undefined as the
waste from one process provides the resources for the next, in
analogy to highly evolved natural ecosystems (type III
ecology).3 Of course, any such system will require energy as
an input but should be optimized to keep energy inputs at a
minimum and optimally maximize the use of renewable energy.
It should be appreciated that the definition of a sustainable
polymer as members of a class of materials that exhibit closed-
loop life cycles (Figure 1, dotted lines) is not universally
accepted; few existing plastics today would meet this definition.
We have adopted this definition as an aspirational goal in the
spirit of sustainable development.
Figure 1. Different models of industrial ecology for plastics industry
and innovation leverage points (ILP) for the development of 1.2. Scope of Review
sustainable polymers with closed-loop life cycles.
This review is framed in the context of the opportunities and
source of modern plastics are typically byproducts of oil and gas challenges for the development of plastic materials that would
refining; extraordinarily sophisticated and efficient catalytic exhibit closed-loop life cycles as part of a future industrial
technologies have been developed for converting these low cost ecology where all resources, products, and “waste” form a
feedstocks to plastics on a vast scale. Nevertheless, once regenerable system sustained only by the input of energy.3,7,9
introduced into the marketplace, the majority of these materials Several excellent reviews on sustainable polymers have been
end up in the environment as waste. While the economic published in the past few years.12−23 The majority focus on the
benefits of this industrial model of resource utilization have use of renewable resources for feedstocks and monomers,
been significant in the past century, the environmental particularly biomass-derived monomers and processes for the
consequences are sobering. More than 300 million metric generation of polymers derived from these mono-
tons of plastic waste is generated every year, with 4.8 to 12.7 mers.12,13,15,16,24−27 Several other reviews discuss the topic
million metric tons entering the ocean, which has negatively from the end-of-use perspective, covering different families of
impacted the marine environment.4 The vast scale of recyclable and biodegradable polymers.20,23,28−31 Herein, we
manufacturing and the environmental consequences associated attempt a system-level analysis, where all aspects of the plastics
with disposal have illuminated the limits to which the planet life-cycle are considered, which include the natural resources
can cope with our current “take, make and dispose” model of that provide the feedstocks and monomers, as well as the means
resources utilization. Furthermore, fossilized carbon remains an of converting these to polymer products, their life cycles, and
abundant but finite natural resource; in the long term, the laws potential for these to exhibit closed-loop life cycles. We have
of supply and demand will lead to increasing costs and volatility framed opportunities and challenges in terms of “innovation
for fossilized natural resources that provide the fuels and the leverage points” (ILP, Figure 1)32 where the development of
attendant feedstocks that form the basis of the current plastics new science and catalytic technologies might lower the
economy. These considerations have stimulated efforts to technological barriers to achieving closed-loop life cycles. To
develop alternative classes of materials that are both more develop a more cyclical model of resource utilization (Figure 1,
sustainable and not so closely tied to petrochemical resources. dotted lines), new economic and technological innovations are
needed at all steps of the plastics life cycle. For reasons we
1.1. Definition describe in more detail below, we focus the review on two
In 1987, the United Nations introduced the concept of classes of materials, polyesters and polycarbonates, as
sustainable development as “addressing the needs of the representative materials to illustrate key principles for the
present without compromising the ability of the future generation of sustainable polymers.
generations to meet their needs”.5 The concept of sustainable Since catalysis represents an essential facet of sustainability
development6 has grown out of an appreciation that our planet and continues to drive the innovation in sustainable polymers,
is small, its natural resources finite. The scale of human we highlight recent advances in the catalytic transformations of
technological activity has had a significant impact on economic renewable feedstocks to monomers, monomers to polyesters
development worldwide, but the attendant environmental and polycarbonates, as well as the chemical and biological
impacts are becoming increasingly apparent. These consid- recycling/degradation of these polymers. Different synthetic
erations have spawned increased attention to sustainability routes and catalytic systems will be compared and discussed in
840 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

the context of sustainability (i.e., atom efficiency, the use of and infrastructure for recovery and reuse are not nearly as well
green catalytic systems, and mild polymerization conditions). developed as those for production of plastics. As described in
1.3. General Principles section 4, new innovations are needed to provide more
technology options for plastics recycling. New strategies are
1.3.1. Feedstocks, Monomers, and Polymers. One needed for the selective depolymerization of plastics either to
guiding principle to sustainable development is that the rate their constituent monomers or other useful intermediates that
of natural resource consumption cannot exceed its rate of could be used as resources for other chemical processes (Figure
regeneration if the system is to be sustainable in the long 1, ILP 5−6). This strategy preserves part of the chemical
term.3,7,9 This has encouraged a focus on renewable resources, complexity and energy43 that was invested to create the plastics.
including those derived from regenerable bio- In order for this process to be energetically efficient, plastic
mass,12,13,15,16,24−27 as sources of feedstocks for the next materials should optimally be derived from polymerization
generation of plastics (Figure 1). For bioderived feedstocks, the reactions that are not strongly exergonic to ensure that, once
rate and time scale of CO2 sequestration can, in principle, be in generated, the materials can be catalytically depolymerized or
balance with the use and release, resulting in an overall chemically recycled with minimal energy inputs (ΔGp° ∼ 0,
“neutral” carbon footprint. This is in contrast to the carbon such that ΔGrxn = ΔGp° + RT ln Q can be optimized for
cycle for petroleum feedstocks, in which the rate and time scale polymerization or depolymerization). This latter criterion
of CO2 sequestration (millions of years) is much slower than motivates our focus on polyesters and polycarbonates.
the use and release (1−10 year time frame).2 Natural polymers Additionally, an effective catalytic system is needed to favor
such as cellulose, hemicellulose, and lignin are among the most the kinetics and the selectivity of the depolymerization reaction
abundant sources of renewably sourced carbon.14,33,34 The to the desired feedstock.
physical properties of these raw materials can be enhanced by Even under the most highly optimized circular materials
physical or chemical modifications, making them useful economies, some fraction of the products will eventually end up
materials for a variety of applications in our daily life.35−37 in the environment. To mitigate the environmental impact of
Small molecules including vegetable oils, terpenes, and CO2 are plastics that end up in the environment, these plastics should be
also useful renewable raw materials.15,16,33 These raw materials biodegradable. This requires engineering of either the polymer
can be further converted to value-added chemicals in integrated composition/formulation or its degradation pathway so that it
biorefineries, in analogy to that currently practiced in petroleum can be broken down by microorganisms to primarily CO2 and
refineries. Biorefinery products such as ethanol, lactic acid, and H2O in a reasonable time frame.20 CO2 and H2O will re-enter
glycerol are important building blocks for sustainable the life cycle through photosynthesis to grow plants that serve
polymers. 14 Among the different biomass sources for as the biofeedstock of sustainable polymers. Although there is
biorefinery, the lignocellulosic feedstock biorefinery38 is no direct correlation between a plastic’s ability to biodegrade
attractive because of the availability of feedstock (e.g., agro- and the resource from which it is derived, many biomass-
food wastes, or even paper wastes) at competitive prices and derived polymers have oxygen-rich backbones and are thus
because the use of these feedstocks does not compete with the more readily degraded than the all-carbon backbones of most
food supply. As the chemical composition of renewable petroleum-based polymers. Such bioderived and biodegradable
resources are quite different from those derived from polymers are therefore emerging as attractive alternatives to
petroleum-based resources, new scientific and technological conventional petroleum-based polymers. However, while
innovations are needed for the development of efficient biodegradable plastics can mitigate some of the environmental
catalytic strategies for converting these resources into platform consequences of plastics that find their way into the
chemicals, monomers, polymers, and products that can be environment, it is only one of the several end-of-life strategies
produced economically on a scale sufficient for commodity to minimize the environmental impact of plastics. It is arguably
materials (Figure 1, ILP1−ILP4). This remains a formidable not the optimal end-of-life option in that the energy invested to
challenge, as discussed in sections 2 and 3. create these materials is wasted and provides benefit only to the
1.3.2. End-of-Use Options. New innovations are also microbes. The end-of-life challenges and opportunities are
needed to develop strategies for the recapture and reutilization described in section 4.
of plastic materials at the end of their useful life. In the United
States in 2014, over 75% of plastics wastes were landfilled,39 1.4. Challenges
with the remainder being either recycled (primarily HDPE and Development of sustainable polymers that meet all the criteria
PET) or incinerated.39,40 However, the amount of landfill space described above is a formidable challenge that will require the
available for discarded plastic wastes is finite. Incineration of concerted effort of academic and industrial scientists and
recovered plastic provides a potential source of energy39 but engineers, industrial ecologists, economists, social scientists,
generates CO2; moreover, rigorous purification and process and policy makers. One of the most significant challenges is
controls are needed to ensure that the hazardous substances are that the plastics of modern society are quite remarkable
not released into the atmosphere.41 Petroleum-based thermo- materials; more sustainable alternatives will have to meet a high
plastics such as PET and HDPE are currently recycled on a bar in terms of the cost and performance in order to be
large scale through a sorting, melting, reprocessing process into economically competitive. Moreover, the economic ecosystem
useful products.42 The current recycling system requires a which has led to our current plastics economy is well-
complicated sorting step which limits the types and amounts of entrenched and economically successful. While considerable
plastic being recycled. Thus, while many petrochemically based research has focused on replacing petroleum raw materials with
polymers are recyclable, the current rates of plastics recycling renewable sources and developing biodegradable or recyclable
are exceedingly low (<10%) and, on a volume basis, are not polymers that minimize waste disposal,16 ultimately the point of
keeping pace with the increased volumes of production.39,42 maximum leverage44 is the socioeconomic system that has led
This is a systems-level challenge, as the economic incentives to our current model of resource utilization. The evolution of a
841 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

socioeconomic system that provides incentives both for 1.5. Metrics Used to Evaluate Sustainability
environmentally sustainable as well as economically sustainable One significant challenge in the field of sustainable polymers is
materials will require systems-level changes at all levels of the lack of a universally recognized standard system or set of
society. Such a system would need to account for the metrics for evaluating and quantifying the sustainability of a
socioeconomic costs associated with the plastics economy, for product or process,7 partially because comprehensive and
example the proliferation of microplastics in the environment,45 quantitative models incorporating all aspects of sustainability
and degradation of natural capital (land, fish stocks, etc.).46 are at an evolutionary stage of development and are not
More specific challenges include the relatively high cost of universally accepted or well-established.7−9 Two concepts that
certain biofeedstocks and biorefinery technologies compared to have informed sustainable polymer design and evaluation are
those that contribute to the current petroleum-based plastics based on metrics such as “Green Design Principles” and “Life
economy. Many commercially viable biorefinery processes such Cycle Assessment”.51
as the fermentation of carbohydrates to carboxylic acids and 1.5.1. Green Design Metrics (GDM). Green Design
alcohols are energy-intensive.14 Additionally, the relatively high metrics embody an approach to assessing the environmental
energy and environmental cost associated with the production, consequences of a product and its associated processes at the
separation, and purification of biomass (e.g., agriculture) can design stage. It is based on green chemistry principles that aim
limit the overall sustainability of the polymer derived from the to reduce the use and generation of hazardous substances in the
biomass. Reductions in cost will require improvements of life cycle of chemical products. The hazards assessed include,
processing technologies, developments of production processes but are not limited to, toxicity, physical hazards (e.g.,
that tolerate lower-purity feedstocks, and the use of waste explosions, fires, etc.), impact on global climate, and resource
agricultural materials (such as corn stover) as feedstocks. depletion.51 The commonly used green design metrics are
Second, sustainable polymers must exhibit comparable or extracted from the well-recognized green chemistry principles
better properties (mechanical and thermal properties, process- such as the “12 Principles of Green Chemistry”, the “12
ability, barrier properties, durability, etc.) than the conventional Additional Principles of Green Chemistry,” and the “12
fossil fuel derived plastics in order for them to be competitive in Principles of Green Engineering”. Specific contents for each
the market. Efforts toward this goal include: (1) synthesizing principle can be found in the literature.43,52,53 Here, we
sustainable counterparts of known petroleum-based plastics condense these principles into themes related to the different
from renewable sources to replicate their performances (e.g., stage of a product’s life cycle: (1) use of renewable and local
polyethylene prepared from bioethylene), (2) developing sources, (2) atom economy, (3) less hazardous reagents and
chemically distinct materials from biofeedstocks with novel synthesis, (4) less waste generation, (5) maximum energy
properties (e.g., self-healing, thermo-reversible) that are not efficiency, (6) products designed for recycling and degradation,
currently available with petroleum-based plastics, (3) chemical and (7) cost efficiency.
and physical modification of natural polymers (e.g., cellulose GDM can be used to guide the design of new materials and
and chitosan grafting). to provide detailed comparison of different aspect of
Third, quantitative metrics of environmental sustainabil- sustainability between materials and processes. One essential
ity7−10 are not nearly as well advanced or universally accepted aspect of GDM is that it evaluates the materials at the most
as metrics of economic sustainability. Life-cycle assessment has fundamental level (i.e., the level of chemical structures and
reactions), which determines the intrinsic properties and
been used to assess the environmental impacts resulting from
environmental outcome of a product. These inherent properties
the production, use, and disposal of a polymer product (see
and associated environmental benefits can propagate through-
section 1.5.2), but the results are variable across studies with
out the various life cycle stages.51 It is for this reason that
quantification being strongly tied to modeling assumptions and
implementation of green chemistry assessment can offer
feedstock choice.47−49 Materials in the early stage of develop- environmental benefits that propagate throughout the life
ment, particularly in academic laboratories, are not routinely cycle. Another essential aspect is that GDM are often used at
examined by those metrics; better communication between the design stage to minimize potential problems that could
scientists and engineers in these traditionally separate fields occur during product development. Therefore, GDM can be
represents an opportunity for more sustainable development. adopted as a preventive metric to identify leverage points for
Another challenge lies in the societal factor of sustainability. new chemistries or processes at the design stage of sustainable
Support from public and government including favorable policy polymers.
and legislation is the key stimulator for technological 1.5.2. Life Cycle Assessment (LCA). Life cycle assessment
innovations on sustainable polymers and for businesses and is a tool to characterize the range and scope of environmental
customers to transition toward more sustainable products. impacts resulting from production, use, and disposal of a
Regardless of which new polymers are selected as alternative product or process. The procedure of LCA is specified by ISO
materials, a significant challenge is to encourage the evolution 14040−14044 series54 as three main phases: life cycle inventory
of a socioeconomic system that enables a high rate of end-of-life analysis, life cycle impact assessment, and life cycle
capture and valorization. To succeed, it is vital for chemists and interpretation. The inventory analysis involves compilation
engineers to work with companies and representatives from the and quantification of the inputs and outputs for a given
public sector to develop viable infrastructures and material product. The data for most petroleum-based polymers and
flows. The New Plastics Economy Initiative, led by the Ellen some biopolymers are available from databases such as
MacArthur Foundation, provides one potential vision of such a Ecoinvent.55 The impact assessment is a process to assess the
system.50 Taken together, these technical, economical, and significance and magnitude of the environmental impact of a
societal challenges explain, in part, why the development of product. The most commonly used standard for assessment is
sustainable polymers is such as a formidable and complex the “Tool for the Reduction and Assessment of Chemical and
challenge. environmental Impacts (TRACI)”56 that assesses the impacts
842 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

by ten categories: acidification, carcinogenic human health aforementioned challenge of incomplete data available on the
hazards, ecotoxicity, eutrophication, global warming potential, environmental fate and costs of discarded materials. The use
noncarcinogenic human health hazards, ozone depletion, phase and disposal phase are too complicated to be included
respiratory effects, smog, and nonrenewable energy use. The because the same polymer can be used in multiple products and
data for these impacts are gathered through individual studies, the disposal scenarios of biopolymers are yet to be studied. In
and some categories are not available for certain polymers, addition, the environmental and human health impacts of the
which makes the comprehensive comparison between polymers byproducts of PLA or PHA biodegradation are not fully
difficult. The findings of life cycle inventory analysis and impact understood. In light of these considerations, the reported
assessment are then combined to reach a conclusion used to rankings should be viewed with some caution; nevertheless, this
inform environmental decisions. study illustrates some of the challenges in assessing the
LCA provides a standardized methodology to compare environmental impacts of plastics and underscores the
different product choices and identifies unintended environ- importance and rich opportunities for developing more
mental trade-offs between impact categories.54 Unlike the green quantitative analyses on the environmental costs and impacts
design metrics which are used at the design stage to reduce of plastics in the environment.
potential problems through avoidance approaches, LCA is often 1.6. Commercial Examples
implemented at a later stage in the product development to
evaluate the actual environmental outcomes associated with In the past decade, a number of synthetic bioderived plastics
different products and processes. That means, LCA results are have been commercialized. One important example is PLA
not available before the production of a product, much less so produced from starch-rich crops such as corn, for example that
during the design of a product as a result of inadequate data on developed by Natureworks LLC under the brand name Ingeo.60
environmental impacts. Although LCA ideally should take into Its physical properties allow it to replace petroleum-based
account all stages of the product life cycle (i.e., cradle-to- plastics in some types of packaging and fiber applications.16,61
cradle), the scope of LCA is often limited to cradle-to-gate due Other commercial examples are poly(butylene succinate) and
to the lack of a complete understanding and inadequate data on poly(butylene succinate adipate) (Bionolle),62 poly(ethylene-
the environmental fate and impact of a polymer. Because of the 2,5-furandicarboxylate) (Avantium’s YXY technology),63 bio-
longevity of hydrocarbon-based plastics in the environment and polyethylene (I’m green by Braskem),64 polyhydroxyalkanoate
the accumulation effects such as degradation of land and (Mirel),65 polyamide 11 (Rilsan),66 and polyamide 410
(EcoPaXX).67 Some commercial polymers are partially renew-
proliferation of microplastics, the long-term impacts of plastic
able, such as poly(trimethylene terephthalate) (Sorona),68 a
wastes are difficult to assess and therefore fall outside of the
polymer alloy of starch and modified poly(vinyl alcohol)
cradle-to-gate analysis. In much of the literature, LCA of
(Mater-Bi),69 and the first-generation PET PlantBottle.70 The
different products and processes often focus only on certain
use of CO 2 as a renewable resource for sustainable
environmental impacts, such as emission of greenhouse gas and
polycarbonates has stimulated several commercialized develop-
nonrenewable-energy use.57,58
ments like QPAC71 and Converge.72 Due to the relatively high
1.5.3. Relation Between the Two Metrics. Since the two
costs and limited material properties, renewable plastics
metrics focus on different aspects of sustainability, the
currently make up only a small share of the global plastics
evaluation based on these metrics can give different results.
market. For example, in 2014, only 1.7 megatons of more than
For example, Landis quantified the adherence to “Green Design
300 megatons of polymers produced globally were bioderived,
Principles” and the environmental impacts (assessed by LCA)
of which the three main products were PLA, biopolyethylene,
of 12 commodity polymers, seven derived from petroleum, four
and bioPET.17 As more advanced and efficient agriculture
derived from renewable sources, and one derived from both.59
processes and synthetic approaches become available, biofeed-
In this analysis, biopolymers such as polylactide (PLA) and
stocks and their conversion to plastic materials should become
polyhydroxyalkanoate (PHA) were ranked the highest based on
more cost-efficient, and we would thus anticipate a larger
green design metrics but in the middle of the LCA rankings
amount of and more versatile renewable plastics coming onto
whereas petroleum polymers such as PE and PP were ranked
the market.
the highest in the LCA rankings but in the middle of green
However, it should be noted that plastics derived from
design rankings. The high ranking of PLA and PHA by the
renewable feedstocks are not necessarily sustainable. Although
green design metrics is mostly due to the relative weight placed
biobased polymers save nonrenewable resources and generally
on the use of renewable materials and their biodegradability/ emit less greenhouse gases for their production, studies have
recycling potential, but their production involves agriculture, shown that they may exert higher environmental impacts in the
fermentation, and multiple chemical processing steps that lead categories of eutrophication and stratospheric ozone deple-
to a low atom economy and an increasing potential for tion.73 Therefore, a thorough analysis of all environmental
greenhouse gas emission and environmental impacts such as impacts of biobased materials in comparison with their fossil
eutrophication, human health impacts, and eco-toxicity. There- fuel-based counterparts is required for strategic decision
fore, one important message from this study is that switching making.
from petroleum polymers to those derived from renewable
resources does not necessarily reduce environmental impacts. 1.7. Opportunities and Challenges for the Development of
Nevertheless, there was a qualitatively positive correlation Sustainable Polyesters and Polycarbonates
between adherence to “Green Design Principles” and a While the sustainability of a polymer should be considered on a
reduction of the environmental impacts within each category systems level that encapsulate a variety of factors in its
(petroleum- or biopolymer). While laudable, this study utilized feedstock, synthesis, and end-of-use options, two important
ranking criteria for the green design metrics that are not criteria are the use of renewable feedstock and the degradability
necessarily widely accepted. In addition, the life cycle analysis or recyclability of the polymer.16 The former reduces the
was limited in scope of “cradle-to-gate”, due to the dependence of polymer production on fossilized carbon sources
843 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Figure 2. Classification of polymers based on feedstock renewability and polymer degradability (adapted from renewability/polymer degradability
matrix defined by Miller74). Traditional polymers are derived from fossil fuels and are nondegradable in nature (lower left quadrant), whereas
sustainable polymers are sourced from renewables and can readily degrade in the environment.

and enables an overall “neutral” carbon footprint on the 2. RENEWABLE MONOMERS


environment,2 and the latter is the key to build a circular Due to the oxygen-rich character of polyesters and poly-
economy that increases resource utilization and reduces waste carbonates, they can be readily sourced from a wide variety of
generation. As summarized in Figure 2, synthetic plastics on the renewable feedstocks. Carbohydrates are the most exploited
market can be classified into four categories according to their and abundant biofeedstock for producing building blocks for
feedstock renewability and polymer degradability.74 Among the these polymers.77 Currently, most of the renewable monomers
families of polymers that can be both renewable and degradable are produced from sugar- or starch-rich crops such as corn and
(upper right quadrant), polyesters and polycarbonates have sugar cane. Transition to nonfood sources such as lignocellu-
attracted the most attention, with a great many of them losic biomass is an important direction to lower the production
costs78−80 and to avoid resource competition with the food
successfully used as substitutes for plastics of petroleum origin.
supply. The two major components of lignocellulose, cellulose
They have met several key principles outlined for a sustainable and hemicellulose, provide an almost unlimited source of C5
polymer (section 1.3), which enables a potentially cyclic model and C6 sugars, provided efficient strategies can be devised to
of resource utilization. First, the monomers of many polyesters recover and purify these feedstocks effectively. Lignin is a
and polycarbonates can be derived from biomass or waste gases complex aromatic polymer that can be broken down to various
via biological or chemical transformations. This allows a faster phenolic monomers. Another common renewable resource is
rate of resource regeneration than resource consumption. plant oil, which is converted to fatty acids and glycerol for
Second, the conversion of these renewable monomers to biodiesel and oleochemical production.16,25,81 Oilseed crops
polyesters and polycarbonates can be achieved via a variety of such as soybean, rapeseed, and oil palm are the major supply of
polymerization methods, which enables diverse backbone oils at present, but algae has turned up as a cheaper alternative
structures and side-chain functionalities that in turn give rise due to its nonfood nature, fast production and recycle rate, and
low use of production inputs.82 In addition to these two major
to versatile physical properties of these polymers. Third, classes of biomass, terpenes and CO2 are also useful feedstocks
because of the ester or carbonate linkages in the main chain and for polyesters and polycarbonates.16,83
the relatively flexible backbones, aliphatic polyesters and The conversion of renewable feedstocks to monomers
polycarbonates can be enzymatically degraded by naturally focuses on two directions: to replace the existing petroleum-
occurring microorganisms such as bacteria, fungi, and algae to derived monomers with their renewable counterparts and to
primarily CO2 and H2O,75,76 which can be converted back to develop new bioderived monomers and polymers with unique
biomass through photosynthesis to complete a cradle-to-cradle structures and properties that current commodity plastics do
life cycle of these materials. Alternatively, these are attractive not have. Identifying the best pathway to convert the
candidates for feedstock recycling as the polymerization aforementioned feedstock to a given monomer is one of the
first challenges for the development of sustainable polymers.
thermodynamics are not as exergonic as those of conventional
The conversion typically involves a mechanical/physical
commodity plastics (PE, PS, and PVC). pretreatment (milling, extraction, and steam reforming),
In the following sections, the catalytic transformations of followed by multiple biological, chemical, or thermochemical
feedstocks to monomers (ILP2), monomers to polymers processes.77 This provides a range of platform chemicals, which
(ILP3), and end-of-use recycling (ILP5, ILP6) or biodegrada- can be directly employed as monomers, or serve as precursors
tion of polyesters and polycarbonates are discussed in detail. for monomer synthesis. The role of catalysis in these
The challenges and opportunities in these steps are highlighted. conversions is critical, and many research efforts are focusing
844 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Figure 3. Structures of diacids, triacids, and hydroxyl acids derived from renewable sources.

Scheme 1. Chemical Conversions of Unsaturated Fatty Acids to α,ω-Diacids, α,ω-Diols, or ω-Hydroxyl Acids

on developing green catalytic processes with high efficiency and succinic acid, muconic acid, or glucaric acid.97 However, none
low cost.84−86 This section intends to give a concise overview of these routes are economically feasible yet.
on the different classes of renewable monomers that have been Fatty acids obtained from transesterification of plant oils have
or can potentially be utilized for the synthesis of sustainable been a long-standing biofeedstock for polymer production.25,98
polyesters and polycarbonates, with a focus on chemical Unsaturated fatty acids containing 1−3 double bonds are
catalysis involved in their productions. More comprehensive particularly attractive substrates since the internal alkene
discussions on this topic can be found in several excellent functionality provides a versatile handle for chemical
review articles.12,33,87−89 modifications (Scheme 1). A variety of long-chain α,ω-diacids
2.1. Diacids, Hydroxyl Acids can be produced through oxidative cleavage of the alkene group
of fatty acids.33 For example, azelaic (C9) and sebacic acid (C10)
Diacids and hydroxyl acids, especially those derived from
are obtained by oxidation of oleic acid and castor oil,
carbohydrates, are arguably the most important chemicals
derived from biomass (Figure 3). They are not only key respectively. Sebacic acid has been used in the production of
monomers for various commodity polyesters and polycarbon- bionylon and polyesters that feature excellent flexibility and
ates but also serve as a platform for a multitude of high-value- chemical resistance. One disadvantage of this transformation is
added products by means of chemical and biological trans- that only part (about half) of the fatty acid chain is utilized in
formations.77 In fact, 10 out of the 15 target biobased building the α,ω-difunctional product and the rest is lost as a coupling
blocks in the 2004 DOE “Top 10” report are carboxylic acids.90 product of less value. Recently, Mecking et al. developed an
1,4-Diacids, including succinic, fumaric, and itaconic acids, elegant approach to incorporate the entire length of the fatty
are some of the most successful platform chemicals obtained acid based on a catalytic isomerization of the internal alkene to
from carbohydrate biorefinery.91 The generation of these the chain end and subsequent ω-functionalization of the
diacids from different natural or engineered microorganisms terminal double bond.99 In one example, they showed that
has been extensively studied and reviewed.88,92−95 Succinic acid isomerizing alkoxycarbonylation of methyl oleate catalyzed by a
is used in the production of biobased PBS; it is also a versatile palladium catalyst bearing phosphorus ligands can convert the
precursor for monomers such as 1,4-butanediol (BDO), γ- internal double bond deep in the hydrocarbon chain to a
butyrolactone (γ-BL), maleic anhydride, and maleic acid.91 terminal ester group with high yield and selectivity.100 Another
Fumaric acid and itaconic acid are useful monomers for the approach based on enzymatic ω-oxidation of the terminal
synthesis of unsaturated polyesters and can be converted to methyl group of fatty acids has been reported for the synthesis
multifunctional monomers such as malic acid and aspartic of α,ω-diacids.101 This requires, however, engineering of yeast
acid.94 Adipic acid, a monomer used for the production of and complex downstream processing to extract the desired
nylon, polyurethane, and polyesters, is one of the most product. These fatty-acid-derived α,ω-difunctional monomers
important diacids in industry and ∼2.5 million tons are provide access to a class of long-chain aliphatic polyesters with
produced annually from petrochemicals.96 Several renewable unique properties that bridge the gap between semicrystalline
routes to adipic acid have been reported, including a microbial polyolefins and the traditional polycondensates.102
synthesis from glucose or fatty acid using engineered Furan-2,5-dicarboxylic acid (FDCA) is a useful biobased
microorganisms, and chemical transformations from bioderived feedstock for the production of aromatic polyesters.103
845 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Poly(ethylene furandicarboxylate) (PEF) produced from industrially produced from lignocellulose on a scale of 400
FDCA is regarded as a renewable substitute for PET.104 kilotons a year, this provides a potentially cheaper route to
Many routes to FDCA have been developed (Scheme 2).105 FDCA.115 A recent strategy based on CO32−-promoted C−H
carboxylation enables an efficient and sustainable synthesis of
Scheme 2. Different Synthetic Routes to FDCA via Either a FDCA.116 The reaction of furoic acid with molten Cs2CO3
HMF or a Furfural Intermediate under a CO2 flow or 8 bar CO2 can give FDCA2− in 70−89%
yield within 12 h. This affords a simple way to transform
inedible biomass and CO2 into a valuable feedstock chemical.
An important low-cost, bioderived triacid is citric acid, which
has been frequently used for the synthesis of biodegradable
thermosetting polyesters for drug delivery, tissue engineering,
and high-tech coating applications.88 Citric acid naturally
occurs in citrus fruits and is industrially produced by sugar
fermentation with a worldwide production of ∼9 million tons
per year. It can use inexpensive carbon sources such as
corncobs and brewery wastes as the feedstock.117
Biomass-derived hydroxy acids provide useful AB-type
monomers for renewable polyesters. Lactic acid is a well-
recognized biobased chemical, currently produced from starch
crops via bacteria fermentation.91 The bioconversion of xylose
The most well-established approach is based on a dehydration to lactate using engineered yeast has also been reported,118
process to convert C6 sugars to hydroxymethylfurfural (HMF), which offers the possibility of using lignocellulose as the
which is then oxidized to FDCA.106 Typically, the two steps are feedstock. Lactic acid is a versatile precursor to a variety of
performed separately, with the dehydration catalyzed by acids monomers, including lactide, propylene glycol, propylene oxide,
and the oxidation carried out under basic conditions using a and acrylic acid.33 A structural isomer of lactic acid, 3-
metal catalyst. The use of lignocelluose as the carbon source hydroxypropanoic acid, can also be obtained from glucose
has been demonstrated,107,108 but fructose is the most common fermentation. It is the monomer for a biodegradable plastic,
source for HMF production.104 The oxidation step can be poly(3-hydroxypropanoic acid),119 as well as a renewable
mediated by a variety of heterogeneous Pt, Au, and Pd catalysts precursor for 1,3-propanediol, malonic acid, and acrylic
with high yield.109 Recently, mild catalytic systems have been acid.33 Another important hydroxyl acid is ricinoleic acid
developed to mitigate the nonproductive degradation of HMF isolated from castor oil. It has been used for the synthesis of
under strongly basic conditions. For example, a MnCo2O4 renewable, degradable elastomers through lipase-catalyzed
supported Ru catalyst was found to afford excellent yield of polycondensation followed by radical cross-linking.27,120 Its
FDCA (99%) from base-free air-oxidation of HMF in water.110
copolymerization with other acids has led to interesting
One-pot dehydration and oxidation of fructose to FDCA is
challenging due to the oxidation of sugar, but recent catalytic copolyesters as tunable delivery systems.121,122 Tartaric acid is
development has improved the efficiency of this process.111 a widely available and cheap hydroxyl diacid present in many
Vanadyl phosphate catalysts can convert fructose to FDCA in fruits and is a principal byproduct of the fermentation process
97% selectivity at 84% conversion;112 Co(acac)3 in a silica sol− in the wine industry. It can also be synthesized from maleic acid
gel system gives 99% selectivity to FDCA from fructose at 72% through epoxidation and hydrolysis.123 Its hydroxy-protected
conversion.113 Alternatively, FDCA can be produced from derivative, 2,3-O-isopropylidine-tartarate, has been used to
furfural through oxidation and carboxylation.114 Since furfural is synthesize polyesters with pendant hydroxyl groups.124

Figure 4. Structures of polyols derived from renewable sources.

846 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Figure 5. Structures of cyclic esters that can be derived from renewable feedstocks.

2.2. Polyols oils. A variety of acids, bases, and lipases can catalyze the
Polyols are another important class of building blocks that can transesterification with high efficiency. In addition, a non-
be derived from renewable resources. The most common bio- catalytic supercritical methanol method has been developed to
derived polyols are summarized in Figure 4. Diols are key solve the two-phase problem of normal methanol/oil
components for linear step-growth polyesters and polycar- mixtures.145 Because of the growing biodiesel industry, glycerol
bonates. Ethylene glycol (EG), a component for PET, is is an attractive platform chemical of biorefineries.91 It can not
traditionally produced from petroleum-based ethylene, via the only be chemically transformed to other monomers but also be
intermediate ethylene oxide. An alternative biobased route is used as alternative substrates to carbohydrates in many
through the conversion of bioethanol to ethylene then to fermentation processes.
ethylene oxide. This process has been used by Coca-Cola in Sugar alcohols such as sorbitol, mannitol, and xylitol are
their PlantBottle packaging since 2009.125 Another promising important polyhydroxy platform chemicals. Sorbitol and
route is the hydrogenolysis of sugar-derived polyols (glycerol, mannitol can be generated from catalytic hydrogenation of
glucose and fructose,146 while xylitol is derived from xylose
xylitol, and sorbitol) catalyzed by supported transition-metal
hydrogenation.147 They can also be produced via a one-pot
catalysts, mostly Ru, Cu, Ni, under basic conditions.126,127 This
chemical transformation of cellulose or hemicellulose combin-
process provides a direct and potentially cheaper route to EG
ing acid-catalyzed hydrolysis and metal-catalyzed hydrogena-
but is complicated by the concurrent generation of other
tion.146,148 However, this process usually results in a mixture of
polyols, like propylene glycol (PG). More selective catalysts are
polyols, and the product distribution is governed by the balance
required in order for this route to become more industrially
between the rates of hydrolysis and hydrogenation, which is
relevant. In addition to hydrogenolysis from glycerol,128,129 PG
catalyst-dependent.126 Sugar fermentations to produce these
can be selectively produced from the reduction of lactic acid, via
polyols are also feasible although it cannot yet compete with
both chemical130−132 and biological catalysis.133 1,3-Propane-
the chemical reduction economically.149−151 These platform
diol (PDO) can be produced via many different microbial
polyols can be further converted to lower alcohols (EG, PG,
fermentation processes from glycerol or glucose.134−136 It has
and glycerol) or to a family of C6 diols 1,4:3,6-dianhydrohex-
been industrially produced on a large scale (over 50000 tons
itols (DAHs) through acid-catalyzed dehydration. Solid acid
annually) by aerobic fermentation of glucose from corn syrup catalysts based on phosphated or sulfated metal oxides (e.g.,
by a genetically modified strain of E. coli, developed by DuPont phosphated Nb 2 O 5 ) have been developed as greener
Tate & Lyle BioProducts.137 1,4-Butanediol (BDO), an alternatives to traditional liquid acids (e.g., H2SO4).146
important component for commodity plastics such as PBS Recently, catalyst-free conditions such as microwave heating
and PBT, is mainly prepared from maleic anhydride by Myriant were also reported.152 Isosorbide can also be synthesized
and Davy Process Technologies. Although maleic anhydride is directly from cellulose via a cascade catalytic process employing
currently produced from petrochemicals, its synthesis via a molten salt hydrate ZnCl2 medium.153 These rigid diols have
catalytic oxidation of sugar-derived furfuraldehyde by hetero- proven particularly useful for the synthesis of polyester and
geneous metal oxides is known.138−141 A leading technology for polycarbonates with high glass transition temperatures.88,154
the synthesis of BDO from renewable resources is through
reduction of biobased succinic acid.142 Alternatively, BDO can 2.3. Cyclic Esters
be directly prepared from sugar fermentation by engineered E. Most cyclic ester monomers employed for the synthesis of
coli.143,144 Longer alkanediols such as 1,5-pentanediol (PTDO) polyesters are not directly available from biomass but can be
and 1,6-hexanediol (HDO) are not directly available from synthesized from platform chemicals in one or multiple steps
biomass but can be produced from hydrogenation of key (Figure 5). The sustainability of these monomers largely
platform chemicals of biorefinery. PTDO is synthesized from depends on the efficiency and environmental impact of the
glutamic acid;33 HDO can be prepared from HMF or adipic catalytic transformations. Despite the successful demonstration
acid. Long-chain α,ω-diols are obtained in a similar way of many bioderived lactones in the laboratory setting, more
through reduction of the corresponding diacids or hydroxyl straightforward and selective synthetic routes need to be
acids derived from plant oils (see section 2.1). developed in the future in order for these processes to become
Glycerol is a triol commonly used for the synthesis of industrially relevant.
branched or cross-linked polyesters and polycarbonates. It is The most successful cyclic ester derived from biomass is
produced by chemical or enzymatic transesterification of plant lactide (LA), the monomer for commodity polylactides (PLAs).
847 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

It is prepared from lactic acid through two steps.155 First, prepared from D-galactono-1,4-lactone in three steps and 47%
lactide acid is converted to PLA oligomers (molecular weight = yield169 or from reduced sugar D-dulcitol in four steps and 30%
1−5 kDa) through polycondensation. The oligomers are then yield.170
selectively depolymerized to LA through an intramolecular Nonstrained five-membered lactones have recently been
transesterification process. This second step requires the use of demonstrated as appealing monomers for fully recyclable
a selective catalyst, typically tin octoate,156 and continuous polyesters. γ-BL is a key downstream chemical of succinic
removal of lactide (e.g., by distillation) to achieve good yield. acid.171 Hydrogenation of succinic acid to γ-BL has been
Both steps are time- and energy-intensive owing to their investigated with a variety of metal catalysts.172 Au/TiO2
reliance on high temperatures and vacuum. Recently, a less catalysts with addition of small amount of Pt can convert
energy-intensive one-step conversion of lactic acid to LA was succinic acid to γ-BL with 97% selectivity at 97% conversion.173
reported using microporous zeolite catalysts.157 The shape- α-Methylene-γ-butyrolactone (MBL) is naturally occurring in
selective property of zeolites enables high selectivity for LA tulip and can also be readily synthesized from sugar-derived
(∼79% at full conversion), devoid of racemization and side- itaconic acid.174 Its derivative, γ-methyl MBL, can be prepared
product formation. Another commonly used six-membered from levulinic acid,175 a biorefinery platform chemical obtained
cyclic ester, glycolide (GA), is prepared in a similar way from from C6 sugars or cellulose.91 These monomers can undergo
glycolic acid via the low molecular weight poly(glycolide) polymerizations at both the alkene and ester site, making them
intermediate. Glycolic acid is naturally occurring in sugar cane potential substitutes for petroleum-based polyacrylates and
or fruits but can also be produced biologically from sugars using polyesters.176,177 Another five-membered lactone, α-Angelica
a genetically modified microorganism158 or chemically from lactone (AL), can also be obtained from levulinic acid through
cellulose using heteromolybdic acid catalysts.159 However, the acid-catalyzed dehydration.33
yields of these processes are typically low, which makes them A few four-membered lactones have been reported to be
hard to compete with the current petroleum-based synthesis. synthetically accessible from renewable feedstocks. β-Propio-
Recently, β-methyl-δ-valerolactone (MVL) has been identified
lactone (PL) is produced from 3-hydroxypropanoic acid by
as a renewable lactone that can be produced from glucose
cyclization.78 β-Malolactonates, a useful class of monomers for
through either a biological or a semisynthetic pathway.160 The
water-soluble poly(malic acid) and amphiphilic PHA-type
biosynthetic route is based on the mevalonate pathway of
architectures, can be synthesized from sugar-derived malic
engineered Escherichia coli strain while in the semisynthetic
route the mevalonate intermediate produced from fermentation acid, aspartic acid, or maleic acid.178 However, the overall yields
is subjected to acid-catalyzed dehydration and Pd/C catalyzed- of all these routes are below 10%.
hydrogenation to yield MVL. This lactone has been used in the Macrolactones are less explored for polymerizations not only
synthesis of thermally recyclable thermoplastics 161 and due to their difficult synthesis but also because of their low
elastomers.162 enthalpies of polymerization179 necessitate polymerization at
Seven-membered lactones including ε-caprolactone (CL) high monomer concentrations in the melt (entropy-driven
and its derivatives can also be generated from biomass through polymerization).180 The most investigated macrolactone for
multistep synthesis. CL is industrially produced by Baeyer− polyester synthesis is pentadecalactone (PDL). It occurs
Villiger oxidation of cyclohexanone. While cyclohexanone is naturally in angelica root essential oil, from which PDL is
currently produced from petro-derived cyclohexane or phenols, isolated at an annual volume of around 1000 t.181 Ambrettolide
its synthesis from lignin-derived aromatic ethers has been is an unsaturated lactone found in the vegetable oil of ambrette
demonstrated using a bromide salt modified Pd/C catalyst in seed.182 Access to larger rings is limited by the availability of
H2O/CH2Cl2 medium.163 Alternatively, CL can be synthesized α,ω-functionalized starting materials. Recently, macrolactones
by oxidative lactonization of HDO. Buntara et al. reported four with ring sizes of 19 and 23 were synthesized via acyloin
routes to HDO from HMF via repetitive hydrogenations using condensation of α,ω-diesters prepared from isomerization
a Rh−Re catalyst on a silica support.164 Several substituted CLs alkoxycarbonylation of oleic and erucic acid, respectively.181
have been synthesized from renewable sources. Racemic CL-1 ROP of these macrolactones generates long-chain polyesters
can be prepared from naturally occurring β-pinene through a with higher and more predicable molecular weights than the
high-yielding (64%) four-step process.165 Disubstituted CL ones prepared from the polycondensation of fatty acids.
derivative CL-2 was generated from the natural product 2.4. Carbonates
(−)-menthol via Baeyer−Villiger oxidation,166 while CL-3 was
derived from carvone, a natural product found in plant oils, via Dialkyl carbonates are considered as green substitutes for
hydrogenation followed by oxone oxidation.167 Given the phosgene for the step-growth synthesis of polycarbonates.
multiple-step nature of the synthesis from not readily available While dialkyl carbonates used to be prepared industrially from
platform chemicals, these monomers are not ideal candidates phosgene, they can now be produced on a large scale using
for sustainable polymers. CO.183 New synthetic technologies that use CO2 as a
Cyclic esters synthesized directly from sugar precursors have renewable carbonyl source are also emerging.184 Dehydrative
been reported.89 Their syntheses typically involve elaborate condensation of CO2 and alcohols is the most straightforward
protecting group chemistry, which compromises the atom and green route to dialkyl carbonates, but the reaction is
economy and production cost of these monomers. Several 1,5- severely limited by the unfavorable equilibrium. An effective
lactones were prepared from the corresponding glycosides, but dehydrating agent such as acetal, orthoester, and a high-
the homopolymerization of these monomers turned out to be performance catalyst, usually a metal alkoxide (e.g., Sn, Ti, and
challenging. One successful example is a sugar lactone (SL-1, Zr) with an acidic cocatalyst [e.g., Sc(OTf)3], are required to
Figure 5) prepared in two steps from D-gluconolactone with drive the equilibrium toward carbonate formation. Dimethyl
90% overall yield,168 which was furthered polymerized using a carbonate (DMC) can be produced by this method in yields of
Sn(OBu)2 catalyst. A 1,6-lactone (SL-2, Figure 5) can be 40−60% based on acetal.185
848 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Scheme 3. Different Synthetic Routes to Six-Membered Cyclic Carbonatesa

a
Only representative or the most recent catalytic and non-catalytic systems are presented.

Figure 6. Cyclic carbonates derived from renewable diols.

Cyclic carbonates are typically synthesized from the triphosgene,187 di-tert-butyl dicarbonate,188 1,1′-carbonyldiimi-
nucleophilic substitution between a diol and a carbonyl source, dazole,189 aromatic carbonates,190,191 and ethyl chlorofor-
traditionally phosgene183,186 or its safer substitutes such as mate.192 Other synthetic routes using CO or ureas as the
849 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

carbonyl source are also known (Scheme 3). Since a variety of significant interest in sourcing epoxides from biomass to
diols can be chemically or microbially prepared from prepare fully renewable polyesters and polycarbonates (Figure
carbohydrates, cyclic carbonates derived from these diols are 7).
partially renewable (Figure 6). PDO and 1,3-butanediol provide
trimethylene carbonate (TMC) and α-methyl trimethylene
carbonate (α-MeTMC). Dihydroxyacetone (DHA) generated
by selective oxidation of glycerol via Pd-catalysis193−195 or
microbial fermentation196 is used to synthesize six-membered
cyclic carbonates 2,2-dimethoxypropylene carbonate
(MeO2DHAC)195−197 and 2,2-ethylenedioxypropane-1,3-diol
carbonate (EOPDC)198 after ketone protection. A series of
seven-membered carbonates (7CC, α-Me7CC, β-Me7CC)
have been prepared from 1,4-diols derived from succinic,
levulinic, or itaconic acid.199 Synthesis of cyclic carbonates
directly from sugar molecule is also possible, but because of the
presence of multiple hydroxyl groups, selective protection is
required. Several cyclic carbonates (Figure 6) prepared by
multistep synthesis from glucose, xylose, or mannose have been
utilized in ring-opening polymerizations.89 Figure 7. Epoxides derived from renewable sources.
The use of CO2 as a safe, cheap, and renewable carbonyl
source for the cyclic carbonate synthesis has been recently
explored (Scheme 3E). The thermodynamics of this Most bioderived epoxides are prepared by catalytic
dehydrative cyclization reaction is unfavorable, similar to the epoxidation of alkene group present in a renewable substrate.
synthesis of dialkyl carbonate from CO2.184 A strategy based on For example, limonene oxide (LO) and α-pinene oxide (APO)
the activation of the hydroxyl group for CO2 addition along are derived from naturally occurring terpenes. Different
with in situ formation of an effective leaving group to facilitate stereoisomers of LO have been synthesized by stereoselective
ring-closure has proven successful. Zhang found that the metal-salen catalysts in the presence of an effective terminal
coupling of 3-chloro-1-propanol with CO2 in the presence of oxidant such as NaOCl or peroxides.204 Williams and Meier
1.1 equiv of Cs2CO3 proceeded efficiently under mild reported the synthesis of various epoxides205 from 1,4-
conditions (40 °C, 1 atm CO2) to provide trimethylene cyclohexadiene (CHD), a common byproduct of olefin
carbonate in 95% yield.200 The use of halide is crucial for the metathesis of polyunsaturated fatty acids such as linoleic
conversion, highlighting the importance of the leaving group. A acid.206 One of the most commonly used epoxides, cyclohexene
slightly different system based on Cs2CO3 was reported by oxide (CHO), can be prepared by selective hydrogenation of
Dyson.201 Instead of starting from haloalcohols, they used alkyl CHD followed by oxidation. Double oxidation of CHD leads to
halides to generate the leaving group in situ and a heterocyclic- bis-epoxides that could be used as cross-linkers. Unsaturated
carbene based catalyst to activate CO2. Using this system, the fatty acids are also common precursors for biobased epoxides.
coupling of substituted 1,2- and 1,3-diols with 1 atm CO2 Methyl 9,10-epoxystearate (MES) derived from oleic acid has
afforded cyclic carbonates in ∼60% yield. A minor pathway been used to prepare polyesters with extremely low Tg (−44
based on the direct coupling of Cs2CO3 with diols and alkyl °C).207 Some epoxy fatty acids are naturally occurring, such as
halide in the absence of CO2 and carbene also contributed to vernolic and coronaric acid from seed oil and 18-hydroxy-9,10-
the formation of cyclic carbonate. Recently, Buchard reported epoxyoctadecanoate (HEOD) from plant suberin.
an efficient two-step procedure using DBU as the base in the Epichlorohydrin (ECH), an important industrial chemical
first step to activate the hydroxyl group toward carbonylation produced on a 1.8 million tons scale annually, can be derived
and NH 3 /TsCl in the second step to facilitate the from glycerin through a two-step reaction.208,209 In the first
cyclization.202 The coupling of diols with CO2 occurred at 25 step, glycerin is dichlorinated by HCl with the assistance of a
°C and ambient pressure, affording cyclic carbonates in 40− carboxylic acid catalyst to generate a mixture of 1,2- and 1,3-
70% yields. This method can be applied to a variety of mono- dichlorohydrin, which is then treated with base to form ECH.
and disubstituted diols and even sterically demanding sugar ECH can be converted to a variety of functional glycidyl ethers,
molecules.202,203 Although these approaches use cheap, safe, which largely increases the structural diversity of the polymers.
and renewable feedstocks and operate under relatively mild One such monomer is furfuryl glycidyl ether prepared from
conditions, the generation of stoichiometric wastes and the ECH and sugar-derived furfural alcohol.210 The furan group
moderate yields of these reactions limit their atomic efficiencies. provides a useful handle for the postpolymerization function-
Future work on improving the efficiency and making the alization via Diels−Alder reaction. A complete survey of
procedure catalytic is required in order for this process to be renewably sourced epoxides and their polymerization reactivity
economically competent to replace the current industrial is available in a recent review by Coates.211
protocols.
3. SYNTHESIS OF POLYESTERS AND
2.5. Epoxides POLYCARBONATES
Epoxides are used as comonomers with cyclic anhydrides or The development of petroleum-based plastics has been driven
CO2 for the synthesis of polyesters and polycarbonates through by the innovations in catalytic petrochemistry and polymer-
ring-opening copolymerization. CO2 is an abundant renewable ization methods. Similarly, the effective uses and trans-
resource, and cyclic anhydride can be readily prepared by formations of renewable feedstocks to sustainable polymers
dehydration of renewably sourced diacids. There is also rely on the development of innovative and green catalytic
850 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Scheme 4. Different Synthetic Routes to Polyesters and Polycarbonates

chemistries. Polyesters and polycarbonates can be synthesized inaccessible through the chain-growth method. SGP also allows
through a variety of routes summarized in Scheme 4. These easy tuning of the physical properties of the polymers by
routes can be divided into two classes: step-growth and chain- varying the structures of building blocks. For example, aliphatic
growth polymerizations. The chain-growth polymerization polyesters with high glass transition temperatures (Tg > 60 °C)
methods include the ring-opening polymerization of a cyclic can be prepared by using a rigid diol such as isosorbide.214 The
monomer and the ring-opening copolymerization of two crystallinity and biodegradation rate of poly(alkylene carbo-
monomers. In addition, a specific class of polyesters, PHAs, nates) can be adjusted by employing diols with different chain
can be synthesized through microbial fermentation. In this lengths.215,216 Moreover, the use of multifunctional monomers
section, different synthetic routes and catalytic systems are such as glycerol and citric acid enables a facile synthesis of
compared and discussed in the context of sustainability (i.e., the biodegradable thermosets.217
use of green catalytic systems, mild polymerization conditions, However, the step-growth nature of SGP leads to several
and high product selectivity). The advantages and limitations of drawbacks, including the difficult control of molecular weights
each method are highlighted. and the inaccessibility to block copolymers or other sequence-
3.1. Step-Growth Polymerization (SGP) specific polymer architectures. In addition, the removal of
byproducts, water or alcohol, from the polycondensation
The traditional way of synthesizing polyesters and polycar- reactions necessitates the use of high temperatures and
bonates is by step-growth polymerization. Polyesters are oftentimes vacuum conditions, which makes this approach
synthesized from polycondensation reactions of diols and very energy-intensive.218 This also precludes the synthesis of
diacids (or diesters) or hydroxyl acids. Polycarbonates are thermally unstable polymers as well as complicates the use of
prepared by polycondensations between diols and a carbonyl volatile monomers.219 Side-reactions and evaporation of
source, traditionally phosgene. Although the step-growth monomers occurring at these harsh conditions can result in a
method has been known since the discovery of synthetic stoichiometric imbalance of reactants, which makes the
polymers in early 20th century,212 it is still the most frequently synthesis of high molecular weight polymers very difficult. To
used method for the synthesis of polyesters and polycarbonates address this problem, a two-step procedure is employed for
in industry. many polycondensation reactions.215,220−222 In the first step,
Due to the wide availability of diols, diacids or hydroxy acids, the esterification or carbonation byproduct (water or alcohol) is
SGP provides easy access to polyesters and polycarbonates with distilled off at ambient pressure to form oligomers. The
diverse backbone structures and side-chain functionalities. reaction between these oligomers is promoted in the second
Given that many diols and diacids can be readily obtained step by slowly reducing the pressure and/or increasing the
from biomass (sections 2.1 and 2.2), SGP provides a temperature. The elimination of the small-molecule byproducts
straightforward route to renewable polyesters and polycar- generated from the chain end coupling drives the increase in
bonates. One such commercial example is biopoly(butylene chain length. This two-stage procedure has enabled the
succinate). Unlike chain-growth polymerizations, SGP is not synthesis of polyesters and polycarbonates with molecular
restricted by the size and substitution pattern of the monomers, weights over 50 kDa.215,223
and it is particularly useful for the synthesis of polyesters and Metal salts such as titanium alkoxides, tin alkoxides, and zinc
polycarbonates with long alkyl segments (>8 carbons) between or magnesium carboxylates are commonly used catalysts for the
each ester/carbonate moiety. For example, SGP of plant oils step-growth synthesis of polyesters and polycarbonates.224−229
derived diols and diacids with >20 methylene sequences These catalytic polycondensations are typically conducted at
generates polyesters with structures and thermal properties very high temperature (>200 °C) and high vacuum to reach
reminiscent of PE. 213 These polyesters are otherwise high conversion. For example, Ti(OiPr)4 can mediate the
851 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Scheme 5. Synthesis of Polycarbonates via Direct Polycondensations of Diols and CO2

polycondensation of several bioderived diesters (succinate, Many biobased carboxylic acids and alcohols contain
adipate, and sebacate) with 1,4-butanediol under high temper- multiple hydroxyl groups. The use of a chemoselective catalyst
ature (up to 240 °C) and in vacuo to generate aliphatic is critical to eliminate the need for tedious and uneconomical
polyesters with Mn > 20 kDa.225,230,231 Melt condensation of ω- protection chemistry239,240 and to enhance the product
hydroxyl fatty acids [e.g., OH(CH2)13COOH], catalyzed by specificity. Takasu found that scandium triflate (Sc(OTf)3)
Ti(OiPr)4 via a two-stage polymerization (200 °C under N2 selectively esterifies primary hydroxyl groups over the
then 220 °C under 0.1 mmHg) generate polyesters with secondary or tertiary alcohol.241 They developed a chemo-
molecular weights ranging from 50 to 110 kDa.232 Meier selective polycondensation of tartaric acid or malic acid with
recently used Sn(Oct)2 to polymerize dimethyl itaconate and polyols such as glycerol and sorbitol under mild conditions [60
long chain diols in the presence of 0.5 wt % of 4- °C, 0.3−3.0 mmHg, 0.5 mol % Sc(OTf)3] to afford linear
methoxyphenol as a radical inhibitor to make linear unsaturated polyesters (Mn between 4 and 26 kDa) with negligible
polyesters with Mn up to 11500 Da.233 This was the first time branching or cross-linking. The pendant hydroxyl groups not
that a high-molecular-weight linear polyester was produced only enable postpolymerization functionalization but also
from itaconic acid without isomerization or cross-linking of the provide a handle to tailor the crystallinity of these aliphatic
vinylic double bond. Further modification of the resulting polyesters.242
unsaturated polyesters by Michael addition yielded polyesters Recently, there has been a surge of interest in developing
with different substituents dangling off the backbone, which organocatalytic processes for the synthesis of polyesters and
provided a means to tune the thermal property of the polycarbonates due to their wide applications in biomedical
materials.233 devices, food packaging, and electronic devices. Several classes
The development of catalytic processes that can operate at of organocatalysts have been used to catalyze the SGP of diols
lower temperatures will reduce the energy input and thus with diacids or carbonates, including strong Brønsted acids such
increase the sustainability of the step-growth synthesis. Metal as H2SO4, triflic acid, p-toluenesulfonic acid (PTSA), bis-
triflates have become a class of popular polycondensation (trifluoromethanesulfonyl)-imide (Tf2NH),243,244 and organic
catalysts for this reason. Because of their strong Lewis acidity, bases such as TBD,221,245−247 MTBD,248 and 1-n-butyl-3-
they can catalyze polycondensation reactions at much milder methylimidazol-2-carboxylate.249 Kobayashi developed a unique
temperatures (<100 °C). Among different metals, bismuth catalytic system250−252 using a surfactant-like acid, such as
triflate [Bi(OTf)3] has received particular attention due to its dodecylbenzenesulfonic acid (DBSA), to mediate the emulsion
low toxicity (Bi3+ is among the least toxic heavy metal ions), polycondensation of aliphatic carboxylic acids and alcohols in
low cost, and stability.234 Using Bi(OTf)3 as the catalyst, the water. Since the water byproduct is expelled out of the
polycondensation of sebacic acid and various α,ω-alkanediols hydrophobic interior of micelles, this reaction can proceed
can be performed at 80 °C under vacuum for 48 h to afford effectively at low temperature (40−80 °C) and 1 atm. This
polyesters with Mn up to 30 kDa.235,236 A few studies using mild and environmentally benign system has been successfully
Bi(OTf)3 to polymerize plant oil-237 and suberin-derived238 applied to several fatty-acid type monomers.253,254 To date, the
long-chain fatty acids generated highly crystalline polyesters activity of organocatalysts for SGP and the molecular weights of
with high melting and degradation temperatures as attractive the resulting polyesters and polycarbonates are not yet
alternatives to petroleum-based polyolefins. comparable to the most efficient metal catalysts. SGP mediated
852 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Scheme 6. Fundamental Steps of ROP of Cyclic Esters (Or Cyclic Carbonates)

by mild and selective enzyme catalysis also provides a metal- variety of diols and dihalides substrates. In another study, Du
free and eco-friendly route to polyesters and polycarbon- employed CH2Cl2 as the dihalide and condensed it with various
ates.255−259 However, the high costs and long reaction times of diols and 1 atm of CO2 to give polymers with Mn up to 11 kDa
enzyme catalysts largely limit their utility in large-scale (Scheme 5e).266 In this case, alternating ether and carbonate
manufacture.257,260 linkages were observed in the polymer. Although the direct use
For the step-growth synthesis of polycarbonates, the of CO2 as a monomer for the synthesis of polycarbonates has
replacement of hazardous phosgene with more renewable and limited success thus far, more effective catalytic systems will
eco-friendly carbonyl reagents is a subject of intensive research. likely enable a more effective synthesis of polycarbonates in the
Organic carbonates such as dimethyl carbonate (DMC) are future.
currently employed as a safer and greener substitute for 3.2. Chain-Growth Polymerization
phosgene. They can be produced on a large scale from CO, or
potentially from CO2.183,184 Recently, several studies have 3.2.1. Ring-Opening Polymerization (ROP). Ring-open-
demonstrated the potential of using CO2 as a renewable and ing polymerization of cyclic esters and carbonates is an efficient
safe carbonyl reagent for the synthesis of polycarbonates method for the synthesis of aliphatic polyester and poly-
(Scheme 5). The direct coupling of CO2 and diols is severely carbonates. It is a type of chain-growth polymerization in which
limited by the reaction equilibrium and requires an effective an initiator is used to ring-open a cyclic monomer to generate a
dehydration system. Kadokawa reported a PR3/CBr4/base reactive center to which further cyclic monomers are added to
system (Scheme 5a) to facilitate the direct polycondensation of form a longer chain (Scheme 6).218 Depending on the nature of
CO2 with alkane diols at room temperature and ambient the initiator, the propagating species can be radical, cationic,
pressure.261,262 Only moderate yields (<60%) and low anionic, or covalent. With the proper choice of an initiating/
molecular weights (Mn up to 6 kDa) were achieved. The catalytic system, ROP of cyclic esters and carbonates can be
polymerization also generated quantitative wastes including performed under mild conditions (room temperature, ambient
CHBr3, R3PO, protonated bases. The first catalytic approach pressure) to give high yield and 100% atom economy, which
for the polycondensation between CO2 and diols was reported makes ROP a highly efficient and green route for the synthesis
by Tomishige in 2016 using CeO2 as the catalyst and 2- of polyesters and polycarbonates. As opposed to step-growth
cyanopyridine as the dehydrating agent (Scheme 5b).263 polymerizations, ROP allows for a fine degree of control over
Quantitative conversions of diols were achieved under 5 MPa several important macromolecular parameters including mo-
CO2, but the molecular weights of the obtained polycarbonates lecular weight, polydispersity, end-group fidelity, regio- and
were extremely low (Mn < 1700 Da). In comparison to the stereoselectivity, and architecture (block, cyclic, star,
direct coupling of CO2 and diols, the use of alkyl dihalides as a etc.).267,268 For this reason, ROP has been the method of
third condensation component has been more successful to choice for the synthesis of tailor-made drug delivery systems
generate polycarbonates with higher molecular weights.264−266 where complex architectures and precise molecular weights are
The first system was reported by Inoue in 1994 (Scheme essential to their performance.269,270 Although cyclic esters and
5c).264 They found that treatment of a mixture of diol, alkyl carbonates are usually not directly available from nature, some
dihalide, and K2CO3 with atmospheric pressure CO2 in an of them can be synthesized from platform chemicals of
aprotic polar solvent generated polycarbonates with 25−90% biorefinery,160,164−167,171,181,271,272 with lactide being the most
yields and Mn of 5−15 kDa (Scheme 5c). Characterizations of successful example.
the resulting polymers suggested an alternating incorporation Because the thermodynamic driving force for ROP is highly
of diols and dihalides into the polycarbonates. This method was dependent on the ring size and substitution pattern of the
recently revisited by Feng and Gnanou, who improved the monomer,273 only a limited set of cyclic esters and carbonates
reaction efficiency by using Cs2CO3 as the base and 10 bar CO2 can be polymerized under practical conditions. Some
(Scheme 5d).265 High yields (>90%) and high molecular representative standard enthalpies and entropies for the ROP
weights (Mn in the range of 10−40 kDa) were achieved for a of cyclic esters with various ring sizes are shown in Table 1.
853 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Table 1. Standard Thermodynamic Parameters for the ROP of Cyclic Esters with Different Ring Sizes273−276
monomer ring size state transitiona ΔH0p (kJ/mol) ΔS0p (J/mol K) [M]eq (mol/L)b
β-propiolactone 4 l→c −82.3 −74 3 × 10−11
γ-butyrolactone 5 s→s −5.4c −39.6c 13.24c
δ-valerolactone 6 s→s −8.4d −14.7d 0.18d
L-lactide 6 s→s −22.9e −41.1e 1.2 × 10−2e
ε-caprolactone 7 s→s −14.0f −6.0f 7.0 × 10−3f
tridecanolide 14 l→l −8 26 2.3 × 10−2g
pentadecanolide 16 l→l 3 23 0.70h
a
State of monomer and polymer: l (liquid), s (solution), and c (semicrystalline solid). bEquilibrium monomer concentration at 298 K. cDetermined
in CD2Cl2, [M]0 = 10 M. dDetermined in dichloromethane/toluene (70:30 v/v), [M]0 = 0.35 M. eDetermined in 1,4-dioxane, [M]0 = 1.0 M.
f
Determined in dichloromethane/toluene (70:30 v/v), [M]0 = 2.0 M. gAt 430 K. hAt 370 K.

More extensive tabulations can be found in the litera- well-defined structures and architectures. In addition, catalytic
ture,273−276 and it should be appreciated that the values are processes performed at mild temperature, low catalyst loading,
sensitive to the standard states specified or the conditions in and solvent-free conditions are particularly attractive from a
which they are measured (solvent, state of monomer/polymer). sustainability standpoint.
For small- and medium-size rings (ring size <8), ring-opening is A variety of initiator/catalyst systems have been developed
driven by enthalpy as a result of release of ring strain. Four-, for ROP of cyclic esters and carbonates.268,270 Alkali metal
six-, and seven-membered rings are favorable for ROP due to complexes such as butyllithium285 and potassium methoxide286
their large ring strains while five-membered lactones such as γ- can mediate anionic ROPs via a nucleophilic or enolate-type
butyrolactone (γ-BL) exhibits a less favorable enthalpy of initiating mechanism. Because of the strong nucleophilic and/
polymerization and is usually considered as nonpolymeriz- or basic character of the anionic initiator/chain ends,
able.277 However, Chen recently found that at low temper- transesterification and epimerization are commonly observed
atures (e.g., −40 °C), the ROP of γ-BL mediated by for these anionic ROPs. The aggregation of alkali metal salts,
La[N(SiMe3)2]3 or tert-Bu-P4 can proceed to high conversions lithium in particular, often complicates the polymerization
(60−90%).278 In contrast, for cyclic esters with large ring sizes, kinetics.287 Strong acids and acylating agents can polymerize
the ΔHp0 of ring-opening is minimal and polymerization cyclic esters and carbonates through a cationic pathway.273 The
becomes entropy-driven as a result of the increase of synthesis of high molecular weight polymers is rather difficult
conformational freedom.180 ROP of a few macrolactones with these catalysts due to the strong tendency of intra-
(ring size = 14, 16) have been successfully applied to prepare molecular transesterification (cyclization) and proton transfer
long-chain polyesters with similar structures but higher and reactions of the cationic propagating species.288 The most
more predictable molecular weights than the ones obtained commonly used catalysts for ROP are transition and post-
from the polycondensation of fatty acids.279−281 transition metal complexes, which mediate the ROP through a
Similar limitations on ring size exist for cyclic carbonates. “coordination−insertion” type mechanism (Scheme 7).289 As
While the ROP of five-membered cyclic carbonates is
enthalpically disfavored, six- and seven-membered cyclic Scheme 7. Typical “Coordination−insertion” Mechanism for
carbonates (6CCs, 7CCs) can undergo facile and controlled Metal-Catalyzed ROP
ROP under mild conditions. The synthetic versatility of 6CCs
allows the incorporation of a wide range of functional groups to
the resulting polycarbonates, which has led to a versatile
material platform for biomedical applications. A comprehensive
list of these functional 6CCs is available in a recent review by
Dove.282 Recently a class of strained 8-membered cyclic
carbonates derived from diethanolamine were prepared and
polymerized to well-defined N-substituted polycarbon-
ates.283,284 The N-substitution not only enhances the ring
strain but also provides a synthetic handle to introduce
functionality to these polycarbonates.
Common to all chain-growth polymerizations, chain transfer
can compete with chain propagation to undermine the control
on the structure and molecular weight distribution of the
resulting polymers.273 As depicted in Scheme 6, intramolecular
chain transfer such as cyclization (kcy) and backbiting (kbb) of
chain end onto an internal ester group leads to the formation of
cyclic polyesters; intermolecular chain transfer (transesterifica- the monomer and alkoxide chain end are both covalently
tion) results in chain scrambling (ksc) and broadens the bound to the metal center, the catalyst can exert a great level of
molecular weight distribution. The relative rates of these steps control on the rate and selectivity of the polymerization. Simple
are mainly controlled by the catalytic processes. Therefore, the metal alkoxides and carboxylates such as tin octoate are cheap
development of selective catalysts that favor the chain and convenient catalysts for ROPs and are used industrially to
propagation over any chain transfer and termination events is produce PLA. However, when it comes to selectivity, especially
critical for the synthesis of polyesters and polycarbonates of stereoselectivity, those well-defined single-site metal catalysts
854 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Figure 8. Representative metal catalysts with ancillary ligands for stereoselective ROP of LA.

with ancillary ligands are more competent. The ancillary ligand stereocomplex) have been successfully synthesized by metal-
has proven to be an important element in ROP catalysis to fine- based catalysts.294,299 Studies of the resulting polymers have
tune the electronic and steric properties of the metal center and shown a strong correlation between the thermal and
to control the selectivity of polymerization. Many review mechanical properties of PLAs and their stereosequences.300
articles268,290−296 have been published in the past decade on the Some representative metal catalysts with high stereoselectivities
synthesis and catalytic performance of different types of metal are shown in Figure 8. Aluminum complexes supported by
catalysts, with Al, Zn, group 3, group 4 metals supported by N- SALEN-type ligands301−309 are some of the most stereo-
or O-type multidendate ligands receiving the most attention. selective catalysts for the ROP of racemic LA, but their low
Some of these catalysts show not only extremely high reactivity reactivities (typical reaction conditions: Al loading >1 mol %,
(TOF > 1000 h−1) but also high regio-/stereoselectivity for the 70−130 °C, and long reaction time) limit their applicability.
ROP of asymmetric monomers. For example, a zinc catalyst Rare-earth metal catalysts supported by phosphasalen ligands
with β-diiminate amido ligand (BDIiPr)Zn[N(SiMe3)2] ex- (such as Ln-phosSALEN in Figure 8) displayed both high
hibits excellent reactivity and selectivity for the ROP of a variety
reactivity and stereoselectivity (Pi up to 0.84 at 25 °C) with the
of lactones and cyclic carbonates.296 With only 20 ppm of this
ability to produce high molecular weight isotactic PLAs (Mn >
catalyst, poly(trimethylene carbonate) with Mn as high as 237
100 kDa) with extremely narrow polydispersities (Đ <
kDa and Đ = 1.68 can be synthesized in quantitative yield
within 2 h.297 No appreciable amount of ether linkage was 1.1).310,311 Other metal catalysts such as zirconium complexes
detected in the resulting polymers, indicating the absence of a with bispyrrolidine-salan ligand312 and zinc complexes with
decarboxylation reaction. In addition, this catalyst showed chiral or achiral auxiliary ligands313−316 (Figure 8) have also
exceptionally high regioselectivity (<1% regio-defects) for the been used to generate iso-enriched stereoblock PLAs from rac-
ROP of methyl-substituted cyclic carbonates.298 LA. Another important bioderived and biodegradable plastics
Given that many cyclic esters and carbonates derived from on the market is poly(3-hydroxyalkanoates) (PHAs). PHAs are
plants contain stereocenters, the stereoselective ROP is an naturally produced by various bacteria (section 3.3) but only
important aspect for the controlled synthesis of sustainable isotactic microstructures are made biologically.317 Syndiotactic
polymers. Among all chiral monomers, the stereoselective ROP poly(3-hydroxyalkanoate)s (PHAs) have been synthesized by
of lactide has been most extensively studied because of the stereospecific ROP of racemic β-lactones by several rare-earth
important application of PLAs. PLAs with different types of metal catalysts.318 This provides new possibilities for engineer-
stereosequences (isotactic, heterotactic, syndiotactic, and ing the properties and applications for these plastics.
855 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Scheme 8. Representative Examples of Using Switchable Metal Catalysts for Monomer Sequence Control319,320,322

Scheme 9. Different Catalytic Mechanisms for the ROP of Cyclic Esters

Recently, the use of switchable metal catalysts to strategically and polycarbonates. Diaconescu reported several redox-switch-
modulate the rate and monomer selectivity of ROP is able group 4 metal complexes supported by ferrocenyl-based
emerging.267 This innovative catalytic strategy has enabled ligand (Scheme 8A).319 They found that the reduced form was
the facile one-pot synthesis of block copolymers of polyesters much more active toward LA than CL while the oxidized form
856 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

had the opposite selectivity. This distinct substrate selectivity The synthesis of well-defined linear polyesters or polycar-
allowed the one-pot copolymerization of the two monomers by bonates with molecular weights over 100 kDa has been difficult
in situ switching of the redox state of the catalyst to synthesize using organocatalysts due to the competitive catalyst-initiated
poly[(LA-minor-CL)-block-(CL-minor-LA)]. In the similar process at low alcohol initiator concentrations. Coulembier
vein, Byers developed a chemoselective copolymerization of developed a zwitterionic ammonium betaine catalyst that
LA and epoxides using a redox-switchable iron catalyst exhibited high selectivity for the ROP of lactide.333 PLA of
(Scheme 8B).320 Unlike the previous case, the two monomers Mn = 97 kDa (DP = 750) with molecular weight conformed to
exhibited orthogonal reactivity with one being completely the monomer-to-alcohol ratios and Mw/Mn = 1.18 can be
inactive during the polymerization of the other. This allowed synthesized by this catalyst. They proposed a charged-assisted
the formation of clean diblock copolymers. Similar copoly- hydrogen-bonding mechanism for this catalyst where it
merization of LA and epoxides were obtained using a redox- activates the alcohol initiator/chain-end through a combination
switchable zirconium complex.321 Williams developed a strategy of hydrogen bonding between the aryloxide group and the
using CO2 as a chemical switch as well as a comonomer for the H(δ+) of ROH and the Coulomb attraction between the
one-pot chemoselective polymerization of epoxide and CL ammonium group and the O(δ-) of ROH (Scheme 10).
(Scheme 8C).322 They found that the dizinc acetate catalyst Despite the weak basicity of aryloxide, these catalysts showed
was an excellent catalyst for the ring-opening copolymerization good reactivity toward L-LA owing to the supplementary ionic
of epoxide and CO2 but not for the ROP of CL. But as soon as activation.
CO2 was removed, the catalyst turned on the ROP of CL with
no incorporation of epoxide. Thus, repetitive introduction and Scheme 10. Proposed Interactions between Ammonium
removal of CO2 resulted in the formation of multiblock Betaine and Alcohol Initiator/Chain End333
poly(carbonate-block-ester). Recently, this approach was
extended to a monomer mixture of epoxide, lactone, anhydride,
and CO2 to synthesize a variety of block copolymers with great
sequence control and predictable compositions in the polymer
chains.323
The replacement of metal catalysts with organic catalysts has
emerged for the synthesis of plastics, especially for those used
in biomedical and microelectronic applications. Since the
pioneering work of Hedrick in 2001,324 the application of Recently, N-heterocyclic olefins (NHOs) have emerged as a
organic catalysts to ROP has witnessed significant progress in class of active and versatile catalysts for the ROP of cyclic esters
the last 15 years and this field has grown to the point that it and carbonates.334,335 These catalysts can function by different
now provides a powerful and, in some cases, better alternative mechanisms: activating the alcohol initiator through hydrogen-
to the use of traditional metal-based catalysts.325,326 Several bonding (Scheme 11A), deprotonating monomers to form
attributes of organic catalysts, including their low cost and wide enolate-type initiator (Scheme 11B), or activating monomer
availability, easy removal from polymers, low toxicity, and through nucleophilic pathway (Scheme 11C). Their catalytic
versatile catalytic mechanisms, make possible a greener and behavior largely depends on the ring size, (un)saturation of the
backbone, and the steric environment of the exocyclic carbon.
more versatile synthesis of polyesters and polycarbonates.
While nonsubstituted NHOs (at the exocyclic carbon, NHO-
Depending on the nature of the catalyst, organocatalysts can
1,3) were ineffective for ROP due to catalyst deactivation
operate through a variety of mechanisms (Scheme 9).326
(Scheme 11C), NHO bearing dimethyl substituents (NHO-2)
Organic bases such as N-heterocyclic carbenes (NHCs) and
proved highly active for the ROP of LA and VL.334 In the
amidines can activate the alcohol initiator/chain ends through presence of an alcohol initiator, the ROP of 50 equiv. L-LA by
hydrogen bonding interactions. Alternatively, they can activate as low as 0.2 mol % NHO-2 reached quantitative conversion in
the monomer through nucleophilic addition. This has led to the less than a minute, resulting in PLAs with low molecular weight
development of zwitterionic ring-opening polymerization for distributions (Đ < 1.2). Under the same conditions, the ROP of
the synthesis of high molecular weight polyesters327 and VL is fast but poorly controlled due to the more significant
polycarbonates284 with cyclic topology. Organic acids such as enolate-based initiation. Chen recently combined NHOs with a
sulfonic or phosphoric acid can catalyze the polymerization Lewis acid Al(C6F5)3 to develop a highly efficient catalytic
through electrophilic activation of monomers. Catalytic systems system.335 The living polymerizations of VL and CL mediated
combining both H-bond donor and H-bond acceptor moieties, by NHO-1/Al(C6F5)3 generated linear polyesters with high
such as thiourea/amine, can simultaneously activate the molecular weights up to 855 kDa (DP = 3200) and narrow
monomer and alcohol initiator/chain end. These systems molecular weight distributions (Đ = 1.03−1.63). End-group
have proven especially selective for chain propagation over analysis revealed an imidazolium enol-like chain end of the final
transesterification, affording narrowly dispersed polyesters and polymers, indicative of an intramolecular proton transfer within
polycarbonates. The diversity in mechanistic pathways not only the zwitterionic intermediates generated from the nucleophilic
provides new opportunities for enhancing the rate and addition of NHOs to the monomer (Scheme 11D). The high
selectivity of polymerization but also enables the synthesis of activity and the ability to operate through different polymer-
polymer architectures that are difficult to access by metal- ization mechanisms make NHO an attractive motif for future
mediated processes. Many review articles have covered different catalyst design and optimization.
aspects of this topic.325−332 Herein, we aim to highlight some Typical nucleophilic organocatalytic ROP is subject to a
recent catalytic developments that feature improved produc- trade-off between rate and selectivity: the most active catalysts
tivity and/or higher level of control on the macromolecular or (i.e., NHCs, TBD) tend to promote side reactions such as
microstructural parameters of polymers. chain transfer and epimerization especially at high conversions
857 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Scheme 11. Structures of NHOs and Their Different Catalytic Pathways for the ROP of Lactones334

while the most selective systems, such as TU/amines, suffer prepared by simple deprotonation of the neutral ureas or
from slow kinetics. Recently a class of (thio)urea anions thioureas. The rate of the polymerization can be tuned over
catalysts were disclosed that exhibit both high activity and several orders of magnitude by modifying the urea or thiourea
selectivity for the ROP of a variety of cyclic esters and substituents. The urea anions are more active than the
carbonates (Scheme 12).321,336 These ionic catalysts are corresponding thiourea anions, with the ability to polymerize
all tested monomers in just seconds (<12s) with excellent
Scheme 12. Family of (Thio)urea Anions for Both Fast and control on molecular weight distributions (Đ = 1.06−1.14).336
Selective ROP DFT calculations suggested a bifunctional mechanism analo-
gous to that of TBD,321 where the (thio)imidiates behave as a
proton shuttle to pick up the proton from the initiator/chain
end, to stabilize the tetrahedral intermediate via double H-
bonds, and to deliver the proton back to the chain end after
ring-opening. The operational simplicity of these urea or
thiourea anions coupled with their high activity, selectivity, and
versatility illustrate the potential of these catalysts for the
sustainable synthesis of polyesters and polycarbonates.
Kiesewetter reported a series of bis- and tris(thio)urea based
H-bond donors (Scheme 13), which exhibited dramatic rate
acceleration compared to the mono-TU/amine systems, yet
retaining the high selectivity for propagation over trans-

Scheme 13. Proposed Monomer Activation Mechanism for bis- and tris-(Thio)ureasa

a
Adapted with permission from ref 338. Copyright 2016 American Chemical Society.

858 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Figure 9. Chiral organocatalysts with high stereoselectivities for the ROP of rac-LA. kL/kD value for the exo-PRO system was not reported.

Figure 10. Representative catalysts for the ROCP of epoxides with CO2 or cyclic anhydrides.

esterification.337,338 The most active one, a tris-urea/MTBD, is membered) that can be incorporated, ROCP is restricted to the
approximately 100 times faster than mono-TU/MTBD for the synthesis of polyesters and polycarbonates whose ester/
ROP of CL, with equally good control on molecular weights carbonate moieties are 2 or 3 carbons apart. On the other
and molecular weight distributions (Đ < 1.05). The rate hand, it is often the only route to polycarbonates whose
acceleration is proposed to arise from an “activated (thio)urea” carbonates are 2 carbons apart since these polymers are
mechanism in which each additional (thio)urea moiety otherwise difficult to obtain from the ROP or step-growth
enhances the acidity via intramolecular H-bond activation method for the reasons mentioned above.
(Scheme 13), supported by DFT calculations. One obvious advantage of epoxide/CO2 copolymerization is
While low stereocontrol at ambient temperature has been a the utilization of CO2 as a monomer. Upcycling of waste gases
long-standing problem for organocatalytic ROP, recent such as CO2 to useful and valuable plastics has long been of
advances in catalyst design have demonstrated the potential interest to researchers, and the ROCP of epoxides and CO2
of organocatalysts to rival the state-of-the-art metal catalysts. represents one of the most successful examples of CO2
Terada and Satoh reported a binaphthol-derived chiral utilization in polymer synthesis.83,344−352 Depending on the
phosphates (BINOL-PP, Figure 9) for the enantioselective epoxide, this method enables 30−50% of a polymer’s mass to
polymerization of rac-LA (kD/kL up to 28) at 75 °C to generate be derived from CO2, with the rest derived from cheap platform
isotactic PLAs (Pm = 0.86) at 49% conversion.339 Chen petrochemicals or potentially from renewable biomass. Because
developed a bifunctional catalyst that combines β-isocupreidine, of the large economic and environmental benefits of this
a previously reported chiral ROP catalyst with low process, it has stimulated a number of commercialization, such
enantioselectivity (kL/kD = 4.4),340 and a chiral thiourea as QPAC by Empower Materials71 and Converge by Saudi
based on the binaphthyl scaffold.341 This catalyst exerts double Aramco.72
stereodifferentiation for the ROP of rac-LA with kL/kD up to 53 The success of ROCP is largely driven by the development of
(for TU-ICD, Figure 9). The ROP of rac-LA mediated by TU- efficient and selective catalysts. Because of the mechanistic
ICD generated isotactic PLAs (Pm = 0.88) at 25 °C and 50% similarities between epoxide/CO2 copolymerization and
conversion. Mecerreyes and Cossió found that a cocatalytic epoxide/cyclic anhydride copolymerization, the two can be
system consisted of a densely substituted proline (PRO, Figure mediated by a similar array of catalysts.343 Early studies focused
9), and DBU was able to kinetically resolve racemic lactide. exo- on heterogeneous catalysts such as zinc glutarate353,354 and
PRO exhibited a strong preference for L-LA, and the ROP of double metal cyanides.355,356 While these catalysts can show
rac-LA by this catalyst at 50% conversion generates a highly good reactivities, they lack sufficient control on the polymer-
isotactic PLAs with Pm up to 0.96 at 25 °C.342 izations and the structures of the active species are ill-defined.
3.2.2. Ring-Opening Copolymerization (ROCP). An The development of single-site metal complexes with ancillary
alternative chain-growth route to aliphatic polyesters and ligands has largely boosted the understanding and the scope of
polycarbonates is the alternating copolymerization of epoxides ROCP.343 In most cases, a nucleophilic cocatalyst is used to
and cyclic anhydrides (for polyesters) or epoxides and CO2 (for increase the activity of the metal complexes, the most common
polycarbonates). While the field of epoxide/CO2 copolymeriza- ones being bis(triphenylphosphine) iminium salts ([PPN]X)
tion is quite mature, with several technologies commercialized, and 4-dimethylaminopyridine (DMAP). Among the different
the progress in epoxide/cyclic anhydride copolymerization is metal catalysts, β-diiminate (BDI) zinc complexes and metal
more recent.211,343 Because of the limitation on the ring size of complexes (Al, Cr, and Co) with Schiff-base ligands (salen and
epoxides (3,4-membered) and cyclic anhydrides (mainly 5,6- their derivatives) are the most successful ones (Figure 10).
859 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Scheme 14. Proposed Mechanism for the ROCP of Epoxides with CO2 or Cyclic Anhydridea

a
Three common side reactions are highlighted in color boxes.

Recently, a series of multinuclear catalysts have been shown to agents are added in large excess compared to the catalyst to
display superior performance than their mononuclear ana- increase the polymerization rate and control the molecular
logues. A comprehensive discussion of these catalysts can be weight of products.360
found in many review articles.83,211,343−345,348,349,351,352 One detrimental side reaction for the ROCP of epoxide/CO2
Although the exact mechanism for each catalytic system may is the formation of five-membered cyclic carbonates (5CCs) as
vary, ROCP of epoxide with CO2 or cyclic anhydride generally a result of intramolecular cyclization (Scheme 14, yellow
follows the coordination−insertion mechanism shown in box).352 Because 5CCs are thermodynamically stable and
Scheme 14.344 The polymerization is initiated by the cannot be reincorporated into the chains, this elimination
nucleophilic attack of a ligand (X) or an exogenous nucleophile causes irreversible loss of yield. The extent of this reaction
onto a coordinated epoxide. This leads to a metal bound depends on the nature of the epoxide, the catalyst, and the
alkoxide to which CO2 (or cyclic anhydride) readily inserts to reaction conditions.83,361 For example, the introduction of
form a metal carbonate (or carboxylate) species. Chain positively charged moieties (e.g., ammonium) to the ligand can
propagation proceeds by the alternative insertion of epoxide suppress the 5CC formation.362−364 ROCP of cyclohexene
and CO2 (or cyclic anhydride) to the metal−oxygen chain end, oxide/CO2 is less susceptible to backbiting than that of
which leads to the desired formation of polycarbonate (or propylene oxide/CO2 due to the unfavorable ring strain of
polyesters) backbones. However, several competitive side cyclohexene carbonate. It is also found that lower temperature
reactions can occur to compromise the selectivity and yield and higher CO2 pressure favor the propagation over
of these polymerizations. The most common side reaction is cyclization.83
the consecutive insertion of epoxide molecules to the metal One appealing aspect of the ROCP approach is the common
alkoxide intermediates, which leads to ether linkages in the availability and structural diversity of epoxide and cyclic
resulting polymer backbone (Scheme 14, green box). This is a anhydride monomers, which can be easily prepared from
problem for many ROCP catalysts since they can also mediate olefins and dicarboxylic acids, respectively. This is in contrast to
the homopolymerization of epoxides. 350 Although this the ROP method, where the preparation of functionalized
polyether formation can be mitigated by using specific metal lactones/cyclic carbonates can be a challenging and tedious
catalysts that are highly selective for copolymerization over process, often requiring multiple-step synthesis. The structures
sequential epoxide enchainment, this is still a prevalent problem of epoxides and cyclic anhydrides that have been employed for
in the ROCP literature. Chain transfer of the propagating ROCP have been recently reviewed by Coates.211 Among them,
chains is another side reaction that occurs in the presence of propylene oxide (PO), cyclohexene oxide (CHO), and phthalic
protic impurities such as water and alcohol (Scheme 14, red anhydride (PA) are the most widely studied. Some epoxides
box). This can cause the molecular weight of the resulting and anhydrides can be prepared from biomass (Figure 7) such
polymers to be lower than that expected.357−359 On the other as limonene oxide (from terpene)365,366 and maleic anhydride
hand, the exploitation of chain transfer phenomenon enables (from furfuraldehyde),138 which are attractive monomers for
immortal polymerization conditions, where the chain transfer fully renewable aliphatic polyesters and polycarbonates. Addi-
860 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Scheme 15. Examples of Terpolymerization of Epoxide, Cyclic Anhydride, and CO2 for the One-Pot Synthesis of Poly(ester-b-
carbonate)

Scheme 16. Overview of Microbial Synthesis of PHAs

tionally, many epoxides and cyclic anhydrides contain two blocks are miscible, only one intermediate Tg higher than
unsaturated moieties that can be utilized for postpolymerization that of the pure polyester and lower than the polycarbonate is
modifications and cross-linking,211 which increases the range of obtained.368 Therefore, the addition of CO2 to an epoxide/
architectures and functionalities that can be achieved through cyclic anhydride copolymerization is a simple way to improve
the ROCP method. the thermal properties of polyesters for many applications that
The large variety of structures yields a diverse set of polymer do not require pure polyesters.
properties that can be easily controlled by the choice of 3.3. Microbial Fermentation
monomer. For example, switching the epoxide from PO to
CHO significantly increases the Tg of polycarbonates from 30− The microbial synthesis by naturally occurring or genetically
modified microorganisms has been practiced industrially for the
40 °C to 80−120 °C.351 By using a different combination of
production of bioplastics.369,370 The polyesters produced by
epoxide and cyclic anhydride, Tg of polyesters can be tuned to
bacteria are polyhydroxyalkanoates (PHAs). PHAs are an
almost any value from −44 to 184 °C.211 A number of studies
important class of biodegradable plastics commonly used for
have demonstrated the effect of backbone structure and length packaging, bottles, cutlery, furniture, biomedical implants, drug
and rigidity of side chains on the thermal and mechanical delivery vehicles, and animal food supplements.371−374 It is
properties of the resulting polyesters and polycarbonates. A among the few plastic materials of which a complete cradle-to-
detailed survey on the structure−property relationship of these cradle life cycle have been established.375
materials can be found in a recent review by Coates.211 PHAs are produced by a variety of microorganisms as a
Copolymerization of different epoxides or anhydrides provides carbon and energy storage material in times of unbalanced
even more possibilities of structures and properties that can be nutrient availability.376 On the basis of the type of carbon
tuned by the ratios of different monomers.83 This leaves many source, the microbial synthesis of PHAs is carried out by either
opportunities for creating polyesters and polycarbonates with of the following two classes of bacteria: heterotrophic bacteria
unique physical properties. (defined broadly as a wide variety of bacteria using multicarbon
Terpolymerization of epoxide, cyclic anhydride, and CO2 is a as their primary growth substrates) that utilize carbon sources
facile strategy for the one-pot synthesis of poly(ester- such as glucose and fatty acids377 and methanotrophic bacteria
carbonate),211,343 a class of biodegradable polymers that have that utilize methane as a primary carbon substrate.378 A general
garnered considerable interest due to their possible superior procedure for PHA biosynthesis is summarized in Scheme 16.
properties than pure polyesters and polycarbonates. Interest- The carbon feedstock is transformed to a common precursor
ingly, cyclic anhydride almost always gets incorporated first acetyl CoA, which is converted to the 3-hydroxyacyl monomer
followed by CO2, leading to the formation of diblock or tapered through key metabolic cycles. The specific pathways of these
block poly(ester-co-carbonates) (Scheme 15). This selectivity is transformations from different carbon sources have been
attributed to the faster rate of anhydride insertion than CO2 reviewed.376,379 The subsequent biotic polymerization of
insertion into the metal-alkoxide intermediate.367 When the monomers by a PHA synthase generates intracellular PHA
861 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Table 2. Different Polymerization Methods for Polyesters


polymerization method advantages drawbacks
polycondensation of diols readily available and renewable monomers; structural diversity of poor control on molecular weight, polydispersity, architecture;
and diacids, or hydroxy polymers energy-intensive reaction conditions
acids
ROP of cyclic esters ability to control molecular weight, polydispersity, microstructure; limited monomer scopes (mostly 4,6,7-membered rings); few
easy access to complex architectures; versatile mechanisms and monomers can be bioderived; difficult to introduce functional
catalysts groups
ROCP of epoxides and easy synthesis of monomers; readily tunable polymer structures and limited to polyesters whose ester moieties are 2 or 3 carbons
cyclic anhydrides properties; ability to control molecular weight, polydispersity, apart; formation of ether linkage
microstructure
microbial synthesis renewable feedstock; bioreactor; low cost and energy input lack of generalizability; difficult control on molecular weights,
polydispersity; extraction of PHAs from bacteria

Table 3. Different Polymerization Methods for Polycarbonates


polymerization method advantages drawbacks
polycondenstaion of diols with readily available diols; renewable monomers poor control on molecular weight, polydispersity,
phosgene/organic carbonate/ architecture; energy-intensive reaction conditions
CO2
ROP of cyclic carbonates ability to control molecular weight, polydispersity, microstructure; limited monomer scopes (mostly 6,7-membered rings);
easy access to complex architectures; versatile mechanisms and synthesis of monomers from renewable resources is not
catalysts trivial
ROCP of epoxide and CO2 use of CO2 as a monomer; ability to control molecular weight, Limited to polycarbonates whose ester moieties are 2 or 3
polydispersity, microstructure; versatile catalysts carbons apart; formation of ether linkage

granules which consist of a PHA core surrounded by a method has its advantages and limitations. Collectively, these
phospholipid layer with attached proteins. The polymer is then methodologies provide a powerful toolbox for the synthesis of
extracted and isolated from the bacteria to provide pure PHA polyesters and polycarbonates.
resin. For certain bacteria under optimized growth condition,
the yield of PHA obtained from the intracellular inclusions can 4. END-OF-USE OPTIONS
be as high as 89% of the organism’s dry weight.380 Recycling is one of the most important end-of-use options to
One important advantage of this biosynthetic route is that reduce the environmental impacts and resources depletion of
the feedstock can be completely sourced from biomass (sugars, plastics. It is a necessary path to realize a close-loop life cycle of
oils) or methane waste. The methanotrophic bacteria that plastics consistent with the principles of circular economy386
produce PHAs can be found at wastewater treatment plants, and industrial ecology,3 where there are no wastes but only
which are often geographically close to sources of waste products. A high level of recycling can potentially allow for
methane.375,381,382 This provides an opportunity for trans- lower energy and material inputs for a given level of products
forming waste gas to valuable plastic materials.383 The most and thus improve eco-efficiency.387 Currently, there are four
commonly occurring PHA through bacteria synthesis is poly(3- main types of recycling processes. (1) Mechanical recycling: on
hydroxybutyrate) (PHB). Depending on the organism and the the basis of mechanical reprocessing of plastic wastes into a
conditions of growth, the molecular weight of PHBs can vary product with equivalent or lower qualities, this is the shortest
from 50 000 Da to over a million Da.384 Methanotrophic path to construct a circular flow of plastics as depicted in
bacteria have been mainly limited to the synthesis of PHB, Scheme 17, but because of the complications in the segregation
although recent work has demonstrated the possibility to
expand the scope of this synthesis through the utilization of Scheme 17. Different Routes of Recycling
cosubstrates.382 In contrast, a much broader substrate scope is
found for heterotrophic bacteria. Collectively, these bacteria
have led to the biosynthesis of over 150 different types of PHA
copolymers.385
Although PHA has been industrially produced through
bacteria fermentation since the 1980’s, there are still several
challenges for the PHA industry.373 First, the production of
PHAs is still costly compared to the traditional fossil fuel-based
commodity plastics. The extraction step that uses a large
quantity of organic solvent is a particularly costly step.
Innovation in the extraction techniques, together with of different types of plastics and the contamination by
improvement in fermentation and the source of feedstock nonplastic materials, only PET and HDPE in the form of
from wastes, are key factors to drive down the manufacturing relatively pure packaging materials are currently recycled
cost of PHAs. Second, the structures and properties of PHAs routinely,388 and the majority of the recyclates is repurposed
currently produced are quite limited. Plenty of opportunities for lower grade materials.389 (2) Chemical recycling: the
exist for the rational engineering of bacteria to produce tailor- selective depolymerization of a plastic into its chemical
made PHAs with more complex structures and functionalities. constituents (monomer, oligomers) provides feedstocks for
To conclude this section, the different polymerization repolymerization to the original products or for the trans-
methods for polyesters and polycarbonates are summarized formation to other chemicals for repurposing. In order for this
and compared in Table 2 and Table 3, respectively. Each process to be energetically feasible, the polymerization
862 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

thermodynamics should not be highly exergonic, to enable done by both automatic and manual methods, which are used
facile depolymerization or feedstock recovery. This approach is to separate PET from other materials such as glass, metals, and
considered uneconomical for petroleum plastics due to the low papers as well as other polymer types. The sorted plastics are
price of petrochemical feedstock compared with the process then ground into flakes and cleaned to remove food residues
costs incurred to depolymerize and recover the monomer.388 and additives. After size reduction, several separation
However, for polymers derived from less exergonic polymer- techniques are further applied to remove impurities from
ization reactions, chemical recycling can provide an attractive PET. The isolated PET is then dried and sold as pellets or
option to recover value, provided that efficient processes can be flakes. The most challenging step is sorting/separation, which
developed and implemented. (3) Biological recycling: the limits the type of polymers that can be recycled and also
degradation of plastics by natural or engineered micro- determines the quality of recyclates. Current techniques of
organisms generates primarily CO2 and H2O (sometimes sorting are based on both the physical properties and chemical
CH4) as end-products, which can return to the life cycle via attributes of the plastics.396 For example, sink/float separation
photosynthesis. This strategy is critical for the biodegradable can effectively separate polyolefins and PS from PVC and PET
plastics that are released into the environment (soil and ocean). due to their different densities. Fourier-transform near-infrared
However, this is the longest path for building a circular flow of spectroscopy (FT-NIR) is the most frequently used technique
the plastics (Scheme 17), and the energy and environmental to determine the polymer type. X-ray detection is particularly
prices paid to build the chemical complexity of monomers are useful to detect PVC, which are 59% chlorine by weight.
completely lost during the degradation processes. Therefore, it Optical sorters such as a color recognition camera are used to
should only be considered when the first two recycling options differentiate between clear and colored PET.388
are not available. (4) Energy recovery via incineration: Since the postconsumer plastics can hardly have the exact
incineration provides a means of capturing the energy value content and purity grade of the virgin materials, the recycled
of waste plastics. However, this process is destructive and does plastics often come with at least some downgrading of
not facilitate a close-loop life-cycle for plastic materials. In properties and consequently less-valuable end products. One
addition, while efficient processes have been developed, cause for the decrease in mechanical properties after recycling is
sophisticated purification and process controls are needed to the contamination of main polymer with other polymers. For
ensure that the hazardous substances are not released into the example, a small amount of PVC contaminant present in a PET
atmosphere.41 Nevertheless, it is the most suitable way for recycle stream will degrade the recycled PET resin owing to
dealing with highly mixed plastics.388 evolution of hydrochloric acid gas from the PVC at a higher
This section is intended to provide an overview on the first temperature required to melt and reprocess PET.388 Addition-
three recycling options for polyesters and polycarbonates. We ally, the ester bond of PET is sensitive to hydrolytic
will focus on the chemical and biological recycling of these degradation. The presence of water could result in chain
polymers since these processes involve different chemical or scission and decrease in molecular weight during the melting
biological catalysis. The nature and mechanism of the catalysts reprocessing.395 Therefore, intensive drying of PET is necessary
or microorganisms used to depolymerize these polymers in a to lower the moisture level prior to melt extrusion.
“controlled” fashion (i.e., that allows for the valorization of the Since the ability to substitute recycled plastic for virgin
obtained synthons or degradation to CO2 and H2O) is polymer depends on the purity of the recovered plastic feed,
emphasized. The biodegradation of aliphatic polyesters and mechanical recycling of rigid plastic bottles consisting of a
polycarbonates for biomedical applications has been reviewed single polymer are much simpler and more economical than
elsewhere390−392 and is out of the scope of this review. those multilayer and multicomponent packages. The key to
4.1. Mechanical Recycling enlarge the scope of mechanical recycling rests on improved
collection schemes and innovative sorting and separation
Mechanical recycling has been applied to packaging materials technologies (i.e., more versatile and reliable detectors, more
such as PET and HDPE as a quick way to integrate them back sophisticated decision and recognition software, and coupled
into the production cycle. In accordance with the National detection and separation instrument).
Association for PET Container Resources (NAPCOR), ∼30.1%
PET plastic bottles were mechanically recycled in 2015 in the 4.2. Chemical Recycling
United States.393 The majority (>70%) of recycled PET Chemical recycling has attracted increasing scientific and
(rPET) is used for production of lower-grade products, mainly commercial attention as an alternative with the potential to
polyester fibers, and only a small fraction (∼10%) is used for absorb and circulate very large amounts of plastic wastes. The
closed-loop bottle-to-bottle recycling.394 This is due to a application of catalysis is a key to the success of chemical
common downgrading problem of mechanical recycling, that is, recycling and is expected to put this technology at the forefront
the recovered plastics tend to lose some quality attributes of the of plastic management for sustainable polymers.
virgin plastics, such as color, clarity, or mechanical proper- 4.2.1. Solvolysis. Solvolysis using water, methanol, and
ties.395 Since closed-loop mechanical recycling requires the glycol has been the most common chemical recycling strategy
polymers to be effectively separated from sources of for step-growth polyesters and polycarbonates.41,397 A number
contaminations (other plastics, metals, pigments, adhesives, of companies have implemented PET depolymerization at pilot
etc.) and stable during reprocessing (grinding and melting), the or commercial level, including Loop Industries,398 Carbios,399
only postconsumer plastic wastes that have routinely been Ioniqa,400 and Gr3n Project.401 Different solvolysis processes of
recycled in a closed-loop fashion are clear PET bottles and PETs and bisphenol A (BPA)-type PCs have been developed to
more recently HDPE bottles.388 recover suitable monomers for the production of new PETs
Mechanical recycling of PET comprises several key steps: and PCs.402,403 However, the industrial applications of this
collection, sorting, size reduction and cleaning, further recycling approach are quite limited due to the higher cost of
separation, and drying.388 Sorting of mingled recyclables are recycled monomers compared to virgin feedstocks. The
863 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Scheme 18. One-Pot Chemical Recycling and Repurposing of PET

Scheme 19. One-Pot Chemical Repurposing of PC

depolymerization is typically catalyzed by acids or bases. For strategy was applied to BPA-PC.416 Postconsumer compact
example, the alkali-catalyzed methanolysis of PC in THF at 40 discs, mainly composed of PC, was depolymerized with
°C provides BPA and dimethyl carbonate in over 95% yield ethylene carbonate and repolymerized with isosorbide and
within 35 min.404 Achilias developed an efficient depolymeriza- succinic acid to generate copolyesters for coating applications.
tion method of PET and PC into their monomers and Recently, a quantitative one-step transformation of PC into
oligomers under microwave irradiation catalyzed by high-value poly(aryl ether sulfone)s in the presence of a
NaOH.405,406 Glycolysis of PET by a variety of metal or carbonate salt was developed as another strategy for
organic catalysts yields bis(2-hydroxyethyl)terephthalester repurposing these plastic wastes.417 This approach uses a
(BHET) as the major product.403,407 Computational chemistry cascade reaction involving the depolymerization of PC by
provides a useful tool to study the depolymerization K2CO3 and formation of reactive phenoxides, which are then
mechanism, which can in turn guide the development of new polycondensed in situ with bis-fluorinated aryl sulfones to form
recycling strategies. In one example of glycolytic depolymeriza- poly(aryl ether sulfone)s (Scheme 19). Computational studies
tion of PET by TBD or DBU, Horn et al.408 showed that the revealed that this reaction proceeds by the nucleophilic attack
rate-determining step for the depolymerization is the ethylene of the metal carbonate on the PC and the evolution of CO2 gas
glycol involved nucleophilic attack. The calculations suggest as the byproduct. Both the metal carbonate and PC are
that these reactions likely involve a bifunctional mechanism in degraded in this process, which leads to the formation of a
which the alcohol reactant activates the carbonyl group of these metal phenoxide salt that is activated for subsequent reaction
polyesters while the organocatalyst activates the reacting with the aryl fluoride. The computational investigation showed
alcohol. The insights derived from this study led to the that stoichiometric amounts of metal carbonate are needed for
development of a variety of amine-based catalytic systems for the reaction to proceed to completion, but insights provided by
PET depolymerization.409 The catalyst-free depolymerization investigating the mechanism also led them to develop a
of PET and PC in supercritical alcohol410,411 or high-pressure catalytic process using 12 mol % K2CO3 with lower reaction
steam412,413 has also proven efficient for recovering high-quality efficiency.
monomers equivalent to the virgin monomers produced from The acid- or base-catalyzed hydrolysis of aliphatic polyester
petroleum. and polycarbonates has also been widely explored, the results of
In addition to monomer recovery, repurposing of waste which have been used to guide the design and application of
plastics into other valuable feedstocks or into value-added these polymers for therapeutic delivery and medical im-
polymeric materials is an attractive option for plastic recycling. plants.418 The overall degradation process is a complicated
Colonna reported a two-step one-pot strategy for converting interplay between the intrinsic degradation kinetics and the
PET to new polyesters using monomers available from diffusive rate of water to the polymer.419,420 The rate of water
renewable sources.414 In the first step, PET was depolymerized diffusion depends on the hydrophobicity of the polymer and its
using isosorbide to oligomers, which were chain extended by shape or surface area while the intrinsic degradation kinetics is
succinic acid in the second step to form a novel copolyester affected by the pH421 and temperature of the medium422 as well
(Scheme 18). Monobutyltin oxide was identified as the most as the molecular weight423 and microstructures of the
active catalyst for both steps and resulted in minimal polymer.424 The degradation of polyesters and polycarbonates
discoloration. The obtained copolyesters demonstrate proper- is in general much faster under basic conditions than acidic
ties suitable for powder coating applications.415 A similar conditions and under acidic conditions than neutral con-
864 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Scheme 20. Different Chain-Scission Mechanisms for the Hydrolysis of PLA Catalyzed by an Acid or Base

Scheme 21. Catalytic Alcoholysis of PLA to Generate Value-Added Lactate-Type Esters

ditions.421 Jung compared the hydrolytic rate of poly(D,L-lactic or diluents for polymeric resins,430 low-temperature lubri-
acid) (PLA), poly(ε-caprolactone) (PCL), and poly(propylene cants,431 and emulsifiers or dispersants in pharmaceutical
carbonate) (PPC) at various pHs (1−13).425 They found that compositions.432 The alcoholysis of PLAs have been achieved
PPC was less susceptible to degradation in acidic medium than under different conditions,433−436 and the efficiency of these
PCL and PLA, with increased degradation rate as PPC ≪ PLA processes is highly condition-dependent. In the early 1990’s,
< PCL. In contrast, PPC was more readily hydrolyzed in basic Du Pont developed a method using 2−3 equiv of alcohol
conditions, with increased degradation rate being PCL ≪ PLA (ROH for R = Me, Et, Bu) per polymer unit at 150−190 °C
< PPC. The mechanism of chain scission has been debated for a with H2SO4 as the catalyst. 69−87% conversion of high-
long time. A random backbone scission and a chain-end molecular-weight PLA to lactates is achieved within 2 h.437,438
scission (“unzipping”) mechanism were proposed for these Leibfarth reported a very efficient way to recycle PLAs and
polymers. Although the two mechanisms operate simulta- PGA to small esters using TBD as the catalyst.439 With only 1.5
neously during hydrolysis, several studies have found that the equiv of alcohol per ester bond, quantitative alcoholysis to small
chain-end scission is significantly faster than the random esters was observed in only 2 min at room temperature with
backbone cleavage.424,426,427 This is attributed to the hydro- nearly complete retention of stereochemistry. The functional
philic nature of the COOH and OH chain ends compared to group tolerance of TBD allows the introduction of different
the hydrophobic backbone. A detailed chain-scission mecha- polymerizable groups (e.g., olefin) to the ester products
nism for PLA has been established based on model studies of (Scheme 21), which provides a number of opportunities in
PLA oligomers.421,424,428 Interestingly, the exact chain-end the production of new polymer materials. The alcoholysis of
scission mechanism depends on the pH of the medium PLA under catalyst-free, solvothermal conditions (4 equiv of
(Scheme 20). While at acidic pH, the hydrolysis proceeds ROH, 200−260 °C) or catalyzed by metal salts (2 equiv of
through a preferential scission of the end groups and spits out ROH, 140−200 °C) were systematically investigated by
one lactic acid unit at a time; the base catalyzed hydrolysis Sobota.440 The presence of catalyst enables milder reaction
proceeds preferentially via a backbiting process from the OH conditions and among the screened metal catalysts, magnesium
chain end leading to the formation of lactide, which is and calcium alkoxides generated in situ from the metallic
eventually hydrolyzed to lactate. These mechanisms are precursors and alcohols gave the best yields (>90%) and
supported by the fact that when the OH chain end is capped negligible epimerization. Compared to the traditional bulk
via acetylation, the degradation rate significantly decreases and production of lactate esters based on the esterification of lactic
the relatively slow random chain scission becomes dominant.428 acid,441,442 these catalytic alcoholyses of PLAs provide a
The conversion of aliphatic polyester wastes to value-added number of advantages, including starting materials derived
small molecules and polymer precursors has been well- from PLA wastes, reduced reaction time and high yield, simple
demonstrated in the alcoholysis of polyesters, commodity purification, and retention of stereochemistry.
PLAs in particular. It provides a strategy for the production of In addition to the production of small-molecule esters,
industrially useful small esters (Scheme 21). For example, alkyl alcoholysis of PLAs by a substoichiometric amount of
lactates, the alcoholysis products of PLAs, are green substitutes multifunctional alcohols or amines can degrade the polymer
for petroleum-derived esters in applications such as solvents429 chains to lower-molecular-weight telechelic polyols that can be
865 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

used for a variety of industrial processes. Plichta showed that 20 ppm, the degradation proceeded mainly via random intra-
the degradation of PLA with less than 10 wt % diols or and intermolecular ester exchange, generating a significant
diamines occurred readily in xylene at reflux or in bulk at 180− amount of diastereomers and cyclic oligomers.448 In contrast,
200 °C in the presence of Sn(Oct)2, affording hydroxyl- for samples with Sn > 480 ppm, the degradation kinetics
terminated PLAs with molecular weights significantly lower indicated an unzipping depolymerization mechanism along with
than the pristine samples.443 The diols or diamines were minor bimolecular transesterification reactions, leading to the
quantitatively incorporated into the degradation products. This predominant production of L-LA.448,449 Acetylation of the OH
strategy provides a facile access to polymers with various chain end significantly increased the polymer stability and
architectures. For example, the use of multihydroxyl com- impeded the thermal recycling, consistent with an unzipping
pounds such as dipentaerythritol results in star polymers with mechanism initiated from the OH chain end.444 Other
six short arms of PLA. When macrodiols such as PEG are used, transesterification catalysts have been used for the depolyme-
a variety of triblock copolymers can be synthesized. However, rization of PLA.450−453 Noda compared the catalytic efficiency
this method exerts little control over the chain length of each of a series of metal alkoxide, enolate, halide, and oxide salts for
block as the insertion of the diols to the PLA is random. In all the monomer recycling of PLLAs at 190−245 °C under 4−5
cases, the products obtained are not pure triblock structures but mmHg. They found that the yield and selectivity for L-LA were
contain some molar fraction of homopolymers of PLA. dependent on the nature of the metal and increased in an order
Nevertheless, this study presents a practical method of chemical of Al < Ti < Zr < Zn < Sn. In the presence of 1 wt % Sn(Oct)2
repurposing of PLA to polymers and oligomers that can be or zinc naphthenate, L-LA was recovered within 2 h in ∼90%
useful precursors for other processes. yield and over 95% selectivity.450
4.2.2. Thermal Recycling (Pyrolysis). While monomer Selective thermal recycling of other thermodynamically stable
recovery cannot be achieved through direct pyrolysis for step- polyesters and polycarbonates has also been reported.454 For
growth polyesters and polycarbonates, this is an attractive example, the thermal depolymerization of PCL and poly(2,2-
recycling option for chain-growth polymers. An emerging dimethyl trimethylene carbonate) can be achieved in bulk at
frontier in sustainable polymers is the design and synthesis of temperature around 250 °C in the presence of a trans-
thermally recyclable polymers that can be depolymerized to esterification catalyst. The use of Bu2Sn(OMe)2 or NaOH
their monomers in high yields, which can then be reused to facilitated a nearly quantitative depolymerization to CL, while
produce virgin quality polymers.23 However, controlled Ti(OiPr)4 gave almost equal amounts of CL and its cyclic
depolymerization through chain-end unzipping (the exact dimer. When the depolymerization of PCL was performed in
reverse process of chain-growth) depends on the thermody- toluene with Bu2Sn(OMe)2 at 110 °C, the products were,
namics of the polymerization: for polymerization reactions that however, a mixture of cyclic oligomers with little formation of
are highly exergonic (ΔG°p ≪ 0) at normal operating CL.455 These studies once again highlight the importance of the
temperatures, the reverse processes are energy intensive choice of catalyst and reaction conditions to achieve selective
under similar conditions. Even so, in many cases the depolymerization to monomer.
depolymerization may be kinetically slow or unselective, The recycling of long chain polyesters derived from step-
requiring the use of a catalyst. Catalysts developed for growth polymerization of ω-hydroxy fatty acids is an important
polymerization can also be used for depolymerization as they aspect of increasing the degree of sustainability of these
reduce the kinetic barrier of both polymerization and renewable polymers. In addition to recovering the linear repeat
depolymerization to the same extent. It will be seen from the units by hydrolysis, the depolymerization of these polymers
examples below that the use of a proper catalyst not only allows through either thermal depolymerization or solution depolyme-
milder depolymerization conditions but also can improve the rization to specific cyclic oligomers provide an attractive option
selectivity of depolymerization reactions. since the resulting cyclic oligomers are difficult to synthesize
For polyesters and polycarbonates with large polymerization through traditional cyclization methods and they can be
enthalpies, a simple pyrolysis process leads to undesired repolymerized to the original polymers through entropy-driven
products. For example, pyrolysis of PLLA at ambient pressure ROP180 in a more controlled fashion than the step-growth
leads to uncontrolled degradation through various nonradical method. Back in 1939, Carothers has described the selective
and radical pathways that yield a mixture of products consisting depolymerization of polydecanoate, polytridecanoate and
of L-LA, meso-LA, cyclic oligomers, acetaldehyde, acrylic acid, polytetradecanoate at 270 °C and 1 mmHg pressure using
CO, and CO2.444−446 To drive the equilibrium toward the MgCl2 or SnCl2 as the catalyst.456,457 Depending on the starting
formation of monomer and suppress other unwanted polyesters, either cyclic monomer or dimer were obtained in
degradation reactions, the thermal recycling is often performed moderate yield (50−70%) and selectivity (50−90%). Various
at lower temperature and reduced pressure to distill the volatile families of macro-lactones have later been synthesized via the
monomer from the polymer melt. The use of a trans- depolymerization of poly(hexamethylene succinate), poly-
esterification catalyst is a key factor in promoting the selective (tetraethylene glycol succinate), poly(ω-hydroxy acids)
depolymerization under milder conditions. Kopinke et al. ([−(CH2)x−1-CO2-]n, x = 8, 10, 11, and 12).458,459 Hodge
compared the thermal degradation of PLAs in the absence and showed that the depolymerization of polyundecanoate in 2%
presence of Sn(Oct)2.446 While the degradation without the w/v chlorobenzene in the presence of a tin catalyst produced
assistance of catalyst was slow and uncontrolled, the presence 90% yield of cyclic oligomers with cyclic dimer being the major
of residual Sn (from the polymerization process) reduced the product (56%). The isolated cyclic dimer can undergo ROP in
depolymerization temperature by 60 °C and increased the neat with a tin catalyst to regenerate polyundecanoates. The
selectivity for lactide recovery. Several groups further studied molecular weights of these polymers can be significantly higher
the effect of the Sn concentration on the depolymerization than the original step-growth polymers.459
efficiency. It was shown that both the depolymerization rate Recently, Chen discovered appropriate thermodynamic and
and selectivity increased with Sn concentration.447−449 For Sn < catalytic conditions to polymerize the low-strain γ-BL into high
866 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

molecular weight poly(γ-BL)s with cyclic or linear topologies ized by a highly active catalyst TBD in bulk at room
(Figure 11).278,460 At −40 °C and ambient pressure, the ROP temperature to obtain a rubbery polyester,461 which can then
be selectively depolymerized back to MVL in 97% yield at 225
°C, 100 mTorr.161 On the basis of this chemistry, they
developed chemically recyclable thermoplastic polyurethanes161
and novel polyester elastomers.162 The former case exploited
the reversibility of the urethane bond to generate the OH end
group to initiate the depolymerization of PMVL block while in
the latter case a small amount of alcohol is intentionally added
to the elastomers to form the OH chain end for recycling. In
both cases, the thermal depolymerization of the materials leads
to complete recovery of MVL. Another monomer, benzodiox-
ipinone (BDP, Figure 11), was utilized by Shaver to establish a
clean and selective polymerization and recycling process using
an aluminum-salen catalyst.462 The resulting polyesters with
Figure 11. Chemical recycling of monomers with low enthalpies of integrated aromatic groups are potential recyclable alternatives
polymerization. for commodity polyesters like PET. These examples clearly
demonstrate the advantage of using monomers with low
of γ-BL proceeded smoothly in the presence of a catalyst polymerization enthalpies for the feedstock-recycling process.
(La[N(SiMe3)2]3 or tert-Bu-P4) to high conversions within 24 However, their exceptional recyclability comes with several
h, generating poly(γ-BL) with Mn up to 30 kDa and Đ ∼ 2.0. intrinsic drawbacks. First, because of the low polymerization
The topologies of the products can be controlled by the enthalpy of the monomer, the polymerization reactions and
La[N(SiMe3)2]3/ROH ratios, with excess ROH favoring the formulation/processing have to be performed at low temper-
formation of linear chains. A clean and quantitative thermal ature (often below room temperature) in order to achieve
recycling of these polymers was achieved by simply heating the decent yields. This increases the energy cost of the products.
bulk material at 220 °C (for the linear polymer) or 300 °C (for Second, the polymers are intrinsically unstable and tend to
the cyclic polymer). Alternatively, the polymer can readily be degrade in the environmental stresses such as heat or moisture.
depolymerized back to the monomer at room temperature in The low thermal stability of these materials could limit their
the presence of a metal or organic catalyst. Chen later extended practical use and shelf life.
this recycling strategy to α-methylene-γ-butyrolactone (MBL), While the closed-loop recycling of PHA can be realized by
which is even more challenging to polymerize than γ-BL due to several biotic pathways,463,464 the chemical recycling of PHAs
the additional complication of controlling the chemoselectivity to the constituent monomers can provide a time-efficient
between a reactive exocyclic olefin and a stable five-membered option.375 Thermal recycling is the most common approach to
lactone.177 Using similar conditions (−60 °C, Ln/ROH = 1/3) recycle PHAs. While uncatalytic pyrolysis performed by heating
to that of γ-BL, they were able to achieve the selective ROP of the polymer to 500 °C under vacuum or N2 leads to the
MBL to produce poly(MBL) with Mn up to 21 kDa, which can formation of PHA oligomers, crotonic acid, and several volatile
be facilely recycled back to the monomer MBL by heating the side products (Scheme 22A),465 catalytic pyrolysis provides a
polymer ≥100 °C for 1 h in the presence of catalysts. These more efficient and selective degradation behavior. For example,
olefin containing polyesters present a class of useful polymer the use of MgO or Mg(OH)2 as the catalyst enables a selective
precursors for cross-linking or thiol-addition to make functional thermal degradation of PHB to trans-crotonic acids (Scheme
thermoplastics and thermosets. Hillmyer and Albertsson have 22B).466 The presence of catalyst not only allowed the pyrolysis
independently measured the thermodynamics of a variety of to be conducted at lower temperature than without a catalyst
substituted lactones and elucidated the effect of the sterics and but also suppressed the formation of oligomeric species. Similar
position of the substituent on the thermodynamics of ring- trend was observed for Ca2+ containing PHB.467 It is proposed
opening.274,275 In general, substituted lactones have relatively that the Lewis acidic nature of Ca2+ and Mg2+ facilitates the
low polymerizability compared to the unsubstituted ones and activation of the six-membered transition state in the unzipping
are thus potential candidates for thermal recycling. Hillmyer depolymerization of PHA (Scheme 23), and thus leads to a
found that β-methyl-δ-valerolactone (MVL) can be polymer- more selective degradation process than in the absence of

Scheme 22. Thermal Degradation of PHBs in the Absence and Presence of a Catalyst

867 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Scheme 23. (A) Chain-End Unzipping Mechanism for Selective Thermal Degradation of PHB and (B) Influence of Mg2+ on the
Unzipping of PHB

catalysts. This catalytic pyrolysis strategy (using Mg(OH)2 as


the catalyst) was used to recycle a variety of short- and
medium-chain length PHAs and further proved the utility of
the recovered monomers as feedstocks for the microbial
synthesis of PHAs.375,381 In this way, a time- and energy-
efficient recycling between monomers (hydroxyalkanoate or
alkenoic acids) and PHAs is established.
4.3. Biological Recycling
Microbial degradation has attracted considerable interest, as
bacteria are responsible for any degradation occurring to
plastics that end up in the environment,75,76 as well as in
industrial composting settings. The first report of microbial
degradation of polyesters dates back to the early 1960’s.
Merrick and Doudoroff et al. noticed that polyhydroxybutyrate
(PHB) “granules” isolated from microbial cell extracts could be
hydrolyzed to β-hydroxybutyric acid while the polymer was
incubated with part of the cell extracts it had been purified
from.468,469 Since then, numerous studies have focused on the
degradation of aliphatic polyesters and polycarbonates by Figure 12. Aliphatic polyesters and polycarbonates amenable to
depolymerization or mineralization by microbial systems.
microbes (mainly bacteria, yeast, and fungi) or cell extracts
from these organisms, in laboratory settings or in the
environment (e.g., marine470 or soil degradation471). The
occurrence of polyesters-degrading microorganisms in the
environment has been classified as follows: PCL = PHB >
PES (polyethylene succinate) > PBS (polybutylene succinate)
> PLA.472 The structure of the aliphatic polyesters and
polycarbonates which have been shown to be susceptible to
biodegradation are depicted in Figure 12. Most experiments
have been performed in aqueous settings (i.e., aqueous
solution, sludge, or soil) with water-insoluble polymers, even
though organic solvents have also been used.
In all cases, enzymes, which are naturally occurring in
microbes (i.e., hydrolases), catalyze the material’s hydrolysis. Figure 13. Classification of the enzymes that trigger the depolymeriza-
These enzymes are either esterases, which hydrolyze esters tion of aliphatic polyesters and polycarbonates and polymers against
bonds, or proteases, which are used by microbes to cleave which activity has been demonstrated.
peptidic bonds. The most commonly studied protease is
proteinase K, which uses serine as a nucleophile. Esterases enzymes exhibit a substrate binding domain (SBD) in addition
relevant to the field of biodegradation of polyesters and to their catalytic domain, which facilitates the adsorption of the
polycarbonates, on the other hand, can be mainly categorized enzyme on the water-insoluble polymers. For the same reason,
into lipases, which cut esters of fats or PHA-depolymerases, the degradation of polyesters and polycarbonates by enzymes
which hydrolyze PHAs. These enzymes were shown to exhibit generally proceeds through surface erosion: the bulk of the
different efficacies depending on the nature of the ester/ material is not affected and the properties (e.g., molar mass,
carbonate bonds and are summarized in Figure 13.471,473 Since crystallinity, etc.) are maintained, while lower molar mass
most aliphatic polyesters and polycarbonates are hydrophobic, fragments are released from the surface.
868 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Microorganisms degrade polymers to use the degradation soil bacteria in aquatic conditions (up to 90% after 60 days for
products as carbon and/or energy source(s). High molar mass PLA and 50% after 40 days for PLLA).477−479 Moreover, faster
polymers are unable to penetrate the membrane of microbial mineralization of PLLA was demonstrated in thermophilic
organisms and so are not as susceptible to microbial attack by conditions because of the close proximity of experimental
intracellular enzymes. Therefore, microbes first express temperatures to the Tg of PLLA.479 PCL and PHB could also
extracellular enzymes, in order to decrease the polymer’s be biodegraded in similar settings (up to 90% degradation after
molar mass and allow for solubilization and assimilation of the 14 days for PHB), while PBS remained mostly intact (3%
oligomers through their cellular membrane. When the chains degradation after 96 days).478 In contrast, Massardier-Nageotte
are short enough to be soluble in the media and cross the et al. reported that their inoculum was inefficient to biodegrade
membrane, further degradation/digestion might take place with both PLA and PCL in anaerobic conditions, evidencing that the
the help of intracellular enzymes. If the degradation stops at the experimental settings, as well as the origin of the inoculum,
monomeric or oligomeric stage, the process is then called have a tremendous impact on the rate of degradation.480
depolymerization, while mineralization is a process that further Overall, anaerobic microbial degradation of aliphatic
transforms the monomeric units into CO2, H2O, and/or CH4 polyesters and polycarbonates remains poorly studied, despite
as end-products. The working mechanism of microorganisms several advantages such as high efficiency, especially for PLA
on hydrophobic polyesters and polycarbonates is summarized and PHAs, and fast rates.
in Figure 14. Noteworthy, when cell extracts are used to 4.3.2. Aerobic Bacteria. Aerobic degradation tests are
depolymerize water-insoluble polymers, only steps 2 and 3 are either performed in composting or aquatic conditions, mostly
observed (see Figure 14). using soil inocula. Because the microorganism has access to O2
in aerobic conditions, the end-products of mineralization are
CO2 and H2O, instead of CO2 and CH4 in anaerobic
conditions. The mineralization of polyesters has been evaluated
using soil (i.e., compost) or diluting the inoculum in growing
medium enriched with minerals and the rates decreased as
follows: PHAs > PCL > PLA.481−484 Similarly to anaerobic
conditions, the nature and provenance of the inoculum, as well
as the experimental setup, considerably influence the
biodegradation’s efficiency. For instance, Enoki et al. isolated
Pseudomonas, Alcanivorax, and Tenacibaculum, which were able
to partially degrade PCL and PHB fibers in deep seawater.
They found that the bacteria, after isolation, were more efficient
in experimental conditions close to their natural occurrence
[i.e., low temperature (4−10 °C) and high pressure (30
MPa)].485
A variety of soil microorganisms have also been isolated and
proved efficient for the mineralization of PLLA on agar plates
Figure 14. Biodegradation of a water-insoluble polymer by a
bacterium (if purified enzymes are used, only steps 2 and 3 are
or in growing medium.486,487 Depending on the experimental
observed). conditions, the isolation of the microorganisms/enzymes that
are thought to be responsible for the polyesters degradation can
be beneficial or detrimental. For instance, the mineralization of
Various test methods have been developed to assess the PHVB was shown to be faster in fertile soil versus isolated
efficacy of the biodegradation, including, just to name a few, enzymes, owing to a more diverse population of degrading
weight loss measurements, quantification of the amount of enzymes in the soil.484 In contrast, Fukushima et al. found that
produced CO2 and/or CH4, analysis of the surface of the the isolation of microorganisms was advantageous to the
polymer after exposure to microorganism/enzyme by micros- mineralization of PLA and PLA nanocomposites.488 Moreover,
copy, and analysis of the supernatant by spectroscopy or the rate of degradation was found to be faster for PLA
chromatography.474,475 The efficiency of microorganisms at nanocomposites than for pure PLA, owing to the presence of
biodegrading plastics have been tested in anaerobic (no O2) or hydroxyl moieties at the surface of nanoclay that can act as
aerobic conditions (i.e., composting), as well as mesophilic nucleophiles.
(20−45 °C) or thermophilic (≈ 55 °C) conditions. Most of Finally, even though most polyesters and polycarbonates are
these experiments have been performed on a lab scale using soil mineralized by microorganisms, Lee et al. showed that
or sludge collected from waste treatment plants. enantiomerically pure (R)-(−)-3-hydroxycarboxylic acids
4.3.1. Anaerobic Bacteria. PLA, PCL, and polyhydrox- could be produced by in vivo depolymerization of P3HB,
yalkanoates (PHAs) have been shown to be susceptible to both in anaerobic and aerobic conditions, in Alcaligenes latus.489
mineralization in anaerobic conditions. The products of They did so by suppressing the intracellular activity of (R)-
anaerobic bacterial digestion are energy, which is used by the (−)-3-hydroxybutyric acid dehydrogenase, which allows
microorganism, and CO2, CH4, and biomass. Kim et al. have metabolizing (R)-(−)-3-hydroxycarboxylic acids to acetoace-
studied the biodegradation of PLA with different mesh sizes tate. A higher yield was observed when the bacterium was
with an initial weight content of PLA versus media of 10 wt % deprived of nutrient and of light in order to block normal
and reported that the production of CO2 and CH4 could go up pathways. This contribution elegantly showed that useful
to 0.21 mmol/day m2 and 0.24 mmol/day m2, respectively, for chemicals that could be reused as monomers or in other
the smaller PLA particles.476 Kunioka et al. and Itävaara et al. chemical industries could be the end-products of microbial
showed that PLA and PLLA, respectively, could be degraded by degradation. Ren et al. later extended this concept to a broader
869 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

range of PHAs, allowing for the synthesis of a variety of R- 4.3.3.2. Lipases. Lipases are the most-studied enzymes for
hydroxyalkanoic acids, the structures of which depend on the the hydrolysis of polyesters and polycarbonates as they degrade
parent PHA.490 a broad spectrum of polymeric structures. Commercially
4.3.3. Cells Extracts. 4.3.3.1. Depolymerases. Merrick and available lipases mainly originate from porcine pancreas,
Doudoroff et al. showed that a mixture of trypsin, a serine Pseudomonas, or the yeast Candida antarctica. Lipase B from
protease, and depolymerase or cells’ extracts from Rhodospir- Candida antarctica was expressed in Aspergilus niger and
illum rubrum composed of an “activator”, “depolymerase”, and immobilized on acrylics beads and is commercialized as
“esterase” could digest the inclusion bodies of Bacillus Novozym 435. Lipases were shown to degrade PCL,504−508
megaterium containing PHB to produce D(−)-β-hydroxybutyric PLA,504,505 PLLA,509 poly(γ-butyrolactone),507 PHB,508,510
ester, with up to 90% hydrolysis after 100 min.469 Since this poly(butylene succinate adipate),511 PCL/PBS blend,511 and
first report, several PHA depolymerases have been isolated on poly(propylene carbonate)508 in PBS at 37 °C, producing
agar plates, mainly from Alcaligenes, Pseudomonas, and mainly short oligomers. The rates of hydrolysis varied
Comamonas, and their structures and properties have been tremendously with the microorganisms/tissue the lipase was
extensively reviewed.491,492 PHA depolymerases usually exhibit extracted from and the polymer.507,512 Overall PCL degrades
a catalytic domain containing an esterase sequence, often the fastest with quantitative depolymerization obtained with
similar to lipase, and a substrate-binding domain (SBD) to Pseudomonas lipase after 4 days in aqueous media, and the
enable adsorption of the enzyme at the surface of hydrophobic addition of PLA or PPC to PCL greatly decreases this rate.
PHAs.493 Several studies have been dedicated to the substrate- Noteworthy, Vert et al. showed that Pseudomonas lipase could
binding domain of PHA depolymerases. For instance, Maeda et break down the crystalline domains of PCL within days in
al. showed by AFM that the adhesive force of the SBD of contrast to hydrolytic degradation that would take several years
PhaZRpiT1 was around 100 pN and that single mutations in the to do the same.504 Finally, the ability of lipases to degrade
SBD to more hydrophobic amino acids allowed for faster rates polycarbonates in buffer has been linked to the glass transition
of hydrolysis.494,495 Moreover, the polymerase’s SBD and temperature of the polymer.513
insoluble PHA are thought to interact through hydrophobic The influence of organic solvents on the degradation of
effects but also molecule-specific contacts.496 Because of this aliphatic polycarbonates and polyesters with lipases has been
mechanism of action, adding a higher amount of enzyme investigated. The addition of organic solvents is advantageous
increases the efficiency of the degradation up to an inhibitory to the solubilization of the polymers but can have a detrimental
concentration that is detrimental to the degradation rates, effect on the stability of the enzyme. Lipases are thought to
only trigger esterification and transesterification reactions in dry
owing to saturation of the substrate that prevents the catalytic
organic solvents, while the presence of water allows for
site from interacting with the polymer.497
hydrolysis.514 In early studies, Kobayashi et al. showed that the
The efficiency of depolymerases was mostly evaluated in Tris
water/organic solvent ratio and the chosen organic solvent had
or PBS buffer solutions with pH ranging from 7.5 to 9.8, to
a strong influence on the degradation products of PCL:
ensure stability of the enzyme. In a systematic study, Doi et al.
macrocycles were obtained in toluene and in dry isopropyl
demonstrated that PHB-depolymerase of Alcaligenes faecalis,
ether, while the addition of ≈0.2% H2O in isopropyl ether led
Pseudomonas stutzeri, and Comamonas acidovorans could only to the formation of linear oligomers.515 After solvent
degrade polyesters with 2 or 3 carbon atoms between each ester evaporation, these oligomers could be repolymerized in bulk
bond and only −H or −CH3 substituents on their backbone using the same lipase catalyst. The same strategy was applied to
[i.e., poly-3-hydroxybutyrate (P3HB), poly-3-hydroxypropio- PBA and polyesters made from a 13-membered lactone with
nate (P3HP), and poly-4-hydroxybutyrate (P4HB)], or AA/BB lower yields. The specific chain cleavage of lipases, as compared
polyesters exhibiting a −O−CH2−CH2−O− motif (i.e., to random scission for uncatalyzed hydrolysis, was demon-
polyethylene succinate and polyethylene adipate).493 The strated. Madras et al. reported the coefficients of enzyme
rates for the best depolymerase/polyester pairs were up to deactivation and of polymer degradation, kd and ks, respectively,
0.28 g/h/cm2. Depending on the provenance of the PHB for PLA, PCL, PGA, and their copolymers using Novozym 435
depolymerase, either monomer and dimer of 3-hydroxybutyric and Lipolase, an engineered lipase from fungus, in several
acid497 or oligomers can be obtained.498−500 Doi et al. also organic solvents and at different water contents.516,517 Acetone
reported that the surface erosion of P3HB and P3HB-co-P4HB allowed for the faster degradation and DMSO was the less
was faster while carrying surface erosion with PHB-depolymer- efficient solvent, evidencing that overall, the depolymerization
ase at 37 °C than by uncatalyzed hydrolytic bulk erosion at 55 rate increased with higher solvent polarity and lower solvent
°C.498 viscosity. PTMC could be depolymerized to TMC in
Several reports mentioned that the purity and crystallinity of acetonitrile at 70 °C with recovery yields up to 80%.518
the PHAs had an influence on the efficacy of the Matsumura et al. investigated the degradation of PHB, PCL,
degradation.469 PHB depolymerase hydrolyses PHB chains and PBA in mixtures of supercritical CO2 (sCO2) and organic
both by exo (chain ends) and endo (middle of the chain) solvents or water with lipase from Candida antarctica. In flow
attacks, as both linear and cyclic PHB could be hydrolyzed by conditions (15 MPa) with 80% sCO2 and 20% toluene at 40
this enzyme. Interestingly, PHB depolymerase can disturb the °C, more than 99% of the products were cyclic oligomers (210
molecular packing of P3HB in crystals, thus facilitating < Mn < 1010 g/mol).519 In contrast, traces of H2O in sCO2 (18
hydrolysis.501,502 Lenz et al. studied the effect of tacticity and MPa) triggered the formation of dicaprolactone (i.e., 1,8-
crystallinity of PHB on the efficiency of PHB depolymerase and dioxacyclotetradecane-2,9-dione) from PCL in yields higher
showed that the rates of degradation were decreasing following than 90%, therefore allowing for a full recycling cycle.520 Using
the trend: atactic> isotactic> syndiotactic. They also concluded the same lipase, Joly et al. demonstrated that PBS could be
that PHB-depolymerases needed at least a [R]-unit diad in transformed to succinic acid (44% yield) by reactive extrusion
order to catalyze hydrolytic cleavage.503 at 120 °C (30 min, 10 wt % lipase).521 Finally, lipases are also
870 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

thought to catalyze the hydrolysis of lipid ester groups in water Catalysis is an enabling science that impacts every step of the
insoluble aggregates like membranes and vesicles.522 plastics life cycle. Many catalytic approaches have been
4.3.3.3. Proteases. Proteases which use serine as the developed to transform renewable raw materials to versatile
nucleophile for the hydrolysis of peptidic bonds, also called monomers and to further convert these monomers to existing
serine proteases, also showed activity for the depolymerization or new polymer materials. The development of renewable
of aliphatic polyesters and polycarbonates. The most potent monomers, the versatile synthetic routes to convert these
one, proteinase K, was first extracted from the fungus monomers to polyesters and polycarbonate, and the different
Tritirachium album and is a broad-spectrum protease known end-of-use options for these polymers are critically reviewed,
to hydrolyze keratin. Proteinase K is stable in the 8−8.6 pH with a focus on recent advances in catalytic transformations that
range and is particularly efficient to degrade PLA.523,524 In lower the technological barriers for developing more
contrast, the depolymerization of poly(glycolic acid),525 sustainable replacements for petroleum-based plastics. Among
PCL,523,524 PHB,523 or poly(trimethylene carbonate)525 with the potentially sustainable alternatives to current commodity
proteinase K is slower than with the aforementioned enzymes. plastics, aliphatic polyesters and aliphatic polycarbonates are
As for most enzymes, the efficiency of proteinase K at highlighted as they can be derived from renewable resources,
degrading semicrystalline polymers increases with decreasing they can be produced by a variety of versatile synthetic
crystallinity contents. For instance, crystalline PLA could not be methods, they exhibit a wide range of useful properties, and
degraded by proteinase K.525,526 However, Wei and Li et al. they can be repurposed to versatile synthetic intermediates or
showed that proteinase K degraded PLLA faster than PDLA, environmentally degraded if they end up in the environment.
mostly owing to higher adsorption of the enzyme at the surface The different catalytic routes to synthesize these polymers from
of PLLA as compared to PDLA.527 In a thorough study on the renewable monomers are surveyed in this review. Significant
influence of the chirality of PLA on the degradation activity of progress has been made in increasing the catalytic activity
proteinase K, they also evidenced that LL, DL, and LD diads (turnover frequency), productivity (turnover number), and
could be cleaved but not DD diads. selectivity toward desired product formation over side reactions
and regio- or stereoselectivity. Recently, catalytic systems that
5. CONCLUSIONS AND OUTLOOK enable the control of polymerization on a more sophisticated
In this review, we describe the challenges and opportunities in level such as architectures and monomer sequences are
the design and development of sustainable polymers that emerging. While metal-based catalysts are in general more
exhibit closed-loop life cycles. These challenges were discussed competent in tackling the aforementioned challenges, organo-
in the context of polyesters and polycarbonates as representa- catalysts have emerged as an efficient alternative. While a
tive materials to illustrate key principles for the generation of variety of active and selective catalysts are available for
sustainable polymers. “Innovation leverage points” (Figure 1)44 converting renewable monomers into well-defined polymers,
were introduced to describe where new science and catalytic data to assess their environmental impacts or sustainability are
technologies might lower the technological barriers to achieving incomplete. Nevertheless, some sustainability metrics and
closed-loop life cycles in a systems-level analysis where all indicators such as green design principles can be used to
aspects of the plastics life-cycle are considered, which include assess different catalytic methods based on their adherence to
the natural resources that provide the feedstocks and green design principles.
monomers, as well as the means of converting these to The end-of-life options of plastics are critical for the
polymer products, their life cycles, and the end-of-use options. development of circular plastics economies, but economic
One of the most formidable challenges for the development disincentives have contributed to the modest rates of recycling
of sustainable polymers is the economic vitality and success of and material recovery. More efficient chemical and biochemical
the current plastics industry, whose development over the past processes are needed to provide better technology options for
half century was spawned by access to relatively cheap energy preserving and recapturing the value of plastics at the end of
and feedstocks derived from petrochemical refining and 60 their useful life. Considerable efforts have been focused on
years of breathtaking scientific and technological develop- biodegradable polymers that can be decomposed naturally or in
ments.528 The evolution of the successful socioeconomic composting sites by microorganisms. The end products are
ecosystem that has led to our current model of resource primarily CO2, H2O, and other metabolites that can re-enter
utilization to one that provides incentives both for environ- the life cycle. Although this end-of-life option mitigates some of
mentally sustainable as well as economically sustainable the issues associated with indiscriminate disposal of plastics in
industrial processes and materials will require systems-level the environment, it is not the only or even best option for
changes at all levels of society, including academic and capturing value. Further developments in the generation of
industrial scientists and engineers, industrial ecologists, polymers that can be selectively depolymerized to the original
economists, social scientists, policy makers, and consumers. monomers or other building blocks are needed to provide more
Critical to that evolution are technology options for more energetically efficient means of preserving the economic and
sustainable alternatives to our current practices. chemical investments made in generating these materials.
Additional challenges for sustainable polymers are ongoing Materials that are designed to be both chemically recyclable and
developments to establish quantifiable goals, indices, and environmentally biodegradable are excellent candidates for
metrics to assess whether a particular activity or plastic is circular material economies and represent the ideal end-of-life
more sustainable than existing processes or materials.7−10 This option for sustainable polymers. For such system to work at
is an important and active area of research and will be critical to scale, we envision the need for an integrated infrastructure that
guide future research and industrial development.3 The takes on waste collection and sorting, depolymerization, and
development of materials that exhibit closed-loop life cycles purification of recovered feedstocks. Similar to mechanical
will require scientific and economic innovations at every stage recycling, plastic sorting and tolerance to contaminants in the
of a plastics life cycle. plastics feeds presents a significant challenge for chemical
871 DOI: 10.1021/acs.chemrev.7b00329
Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

recycling. Ideally, mixed chemical recycling facilities that are Grubbs. Following a year of postdoctoral research with the late
able to handle multiple depolymerizations from mixed materials Professor Pino at the ETH in Zurich, he joined the faculty at Stanford
flows are preferable. University in 1988, where he is now the Robert Eckles Swain Professor
The development of more sustainable alternatives to our of Chemistry. His research interests are in homogeneous catalysis and
current model of linear resource utilization is a crucial but polymer chemistry.
formidable challenge for the development of sustainable
industrial processes, including those involving the production,
use, and recycling of plastics. The innovations over the past 60 ACKNOWLEDGMENTS
years that have provided the economic engine for our current We thank Professor Geoffrey W. Coates and Dr. Scott Allen for
plastics economies have been breathtaking. Future innovations their insights and discussions in framing the content of this
are needed to ensure that the materials of the next century will review. We also acknowledge the Ellen MacArthur Foundation
not only be economically sustainable, but will be sourced, for their constructive comments in revising the manuscript. The
produced, utilized, and repurposed such that they can address authors thank the National Science Foundation (GOALI CHE-
the needs of current and future generations. 1607092) and the Department of Energy (DE SC0005430) for
financial support. X.Z. acknowledges a Stanford Graduate
AUTHOR INFORMATION Fellowship and a LAM Graduate Fellowship.
Corresponding Author
*E-mail: waymouth@stanford.edu. REFERENCES
ORCID (1) Data from PlasticsEurope, Plastics: The Facts 2015. http://www.
plasticseurope.org/Document/plastics---the-facts-2015.aspx (accessed
Xiangyi Zhang: 0000-0003-4290-1600 August 5, 2017).
Mareva Fevre: 0000-0001-6460-9227 (2) Narayan, R. Carbon footprint of bioplastics using biocarbon
Robert M. Waymouth: 0000-0001-9862-9509 content analysis and life-cycle assessment. MRS Bull. 2011, 36, 716−
Notes 721.
(3) Jelinski, L. W.; Graedel, T. E.; Laudise, R. A.; Mccall, D. W.;
The authors declare no competing financial interest. Patel, C. K. N. Industrial Ecology - Concepts and Approaches. Proc.
Natl. Acad. Sci. U. S. A. 1992, 89, 793−797.
Biographies (4) Jambeck, J. R.; Geyer, R.; Wilcox, C.; Siegler, T. R.; Perryman,
Xiangyi Zhang received her Bachelor’s Degree in 2008 from Nanjing M.; Andrady, A.; Narayan, R.; Law, K. L. Plastic waste inputs from land
University, China, where she did her undergraduate research with into the ocean. Science 2015, 347, 768−771.
Professor Jing-lin Zuo on electroactive diruthenium complexes for (5) Report of the World Commission on Environment and Development:
molecular electronics applications. She obtained her Ph.D. in Our Common Future, United Nations General Assembly, 1987.
(6) Holdren, J. P. Presidential address - Science and technology for
chemistry from Stanford University in 2017 under the supervision of
sustainable well-being. Science 2008, 319, 424−434.
Professor Robert M. Waymouth. Her doctoral research focuses on the (7) Hay, L.; Duffy, A.; Whitfield, R. I. The Sustainability Cycle and
development of organocatalytic approaches for the synthesis of Loop: Models for a more unified understanding of sustainability. J.
functional polyesters and polycarbonates as sustainable alternatives Environ. Manage. 2014, 133, 232−257.
to petroleum-based plastics. She is currently working as a Senior (8) Mayer, A. L. Strengths and weaknesses of common sustainability
Chemist at the Dow Chemical Company in Midland, MI. indices for multidimensional systems. Environ. Int. 2008, 34, 277−291.
Mareva Fevre received her Ph.D. in Chemistry 2012 from the (9) Orecchini, F. A. ″measurable’’ definition of sustainable develop-
ment based on closed cycles of resources and its application to energy
University of Bordeaux, France, working on organocatalyzed polymer-
systems. Sustain. Sci. 2007, 2, 245−252.
ization reactions and the precise synthesis of polymeric ionic liquids,
(10) Kajikawa, Y. Research core and framework of sustainability
under the supervision of Prof. Daniel Taton. In 2012, she moved to a science. Sustain. Sci. 2008, 3, 215−239.
postdoctoral position at Duke University to study elastin-like (11) Towards the Circular Economy, Vol I−III, Ellen MacArthur
polypeptides-containing recombinant proteins for anticancer applica- Foundation, 2012−2014.
tions. Since joining IBM Almaden Research Center in 2014, she’s been (12) Gandini, A.; Lacerda, T. M. From monomers to polymers from
part of some of IBM’s programs leveraging polymer science for the renewable resources: Recent advances. Prog. Polym. Sci. 2015, 48, 1−
preparation of high performance materials, as well as for biomedical 39.
applications, including the design of new antimicrobial and antiviral (13) Mathers, R. T. How Well Can Renewable Resources Mimic
materials. Commodity Monomers and Polymers? J. Polym. Sci., Part A: Polym.
Chem. 2012, 50, 1−15.
Gavin O. Jones is a research staff member in the Computational (14) Ragauskas, A. J.; Williams, C. K.; Davison, B. H.; Britovsek, G.;
Chemistry and Materials Research group at IBM Research−Almaden. Cairney, J.; Eckert, C. A.; Frederick, W. J.; Hallett, J. P.; Leak, D. J.;
He earned his Bachelor’s degree at Bard College and completed his Liotta, C. L.; et al. The path forward for biofuels and biomaterials.
Ph.D. in theoretical/computational organic chemistry at the University Science 2006, 311, 484−489.
of California Los Angeles with Prof. Kendall Houk. He performed (15) Yao, K. J.; Tang, C. B. Controlled Polymerization of Next-
postdoctoral research with Prof. Stephen Buchwald at the Generation Renewable Monomers and Beyond. Macromolecules 2013,
Massachusetts Institute of Technology. He has interests in 46, 1689−1712.
mechanisms, catalysis, molecular properties, and polymer formation (16) Zhu, Y.; Romain, C.; Williams, C. K. Sustainable polymers from
and degradation. He was awarded with Foreign Policy Magazine’s renewable resources. Nature 2016, 540, 354−362.
Global Thinker Award (Innovator Category) in 2016. (17) Shen, L.; Worrell, E.; Patel, M. Present and future development
in plastics from biomass. Biofuels, Bioprod. Biorefin. 2010, 4, 25−40.
Robert M. Waymouth received bachelor’s degrees in mathematics and (18) Schroder, K.; Matyjaszewski, K.; Noonan, K. J. T.; Mathers, R.
chemistry from Washington and Lee University and his Ph.D. from the T. Towards sustainable polymer chemistry with homogeneous metal-
California Institute of Technology in 1987 with Professor Robert H. based catalysts. Green Chem. 2014, 16, 1673−1686.

872 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

(19) Vilela, C.; Sousa, A. F.; Fonseca, A. C.; Serra, A. C.; Coelho, J. F. (44) Meadows, D. Leverage Points. Places to Intervene in a System
J.; Freire, C. S. R.; Silvestre, A. J. D. The quest for sustainable http://drbalcom.pbworks.com/w/file/fetch/35173014/Leverage_
polyesters - insights into the future. Polym. Chem. 2014, 5, 3119−3141. Points.pdf (accessed May 11, 2017).
(20) Tschan, M. J. L.; Brule, E.; Haquette, P.; Thomas, C. M. (45) Galloway, T. S.; Lewis, C. N. Marine microplastics spell big
Synthesis of biodegradable polymers from renewable resources. Polym. problems for future generations. Proc. Natl. Acad. Sci. U. S. A. 2016,
Chem. 2012, 3, 836−851. 113, 2331−2333.
(21) Miller, S. A. Sustainable Polymers: Opportunities for the Next (46) Valuing Plastic: The Business Case for Measuring, Managing
Decade. ACS Macro Lett. 2013, 2, 550−554. and Disclosing Plastic Use in the Consumer Goods Industry, 2014,
(22) Schneiderman, D. K.; Hillmyer, M. A. 50th Anniversary United Nations Environment Programme (UNEP). http://hdl.handle.
Perspective: There Is a Great Future in Sustainable Polymers. net/20.500.11822/9238 (accessed August 5, 2017).
Macromolecules 2017, 50, 3733−3749. (47) Posen, I. D.; Griffin, W. M.; Matthews, H. S.; Azevedo, I. L.
(23) Hong, M.; Chen, E. Y. X. Chemically recyclable polymers: a Changing the Renewable Fuel Standard to a Renewable Material
circular economy approach to sustainability. Green Chem. 2017, 19, Standard: Bioethylene Case Study. Environ. Sci. Technol. 2015, 49, 93−
3692−3706. 102.
(24) Gandini, A.; Lacerda, T. M.; Carvalho, A. J. F.; Trovatti, E. (48) Hottle, T. A.; Bilec, M. M.; Landis, A. E. Sustainability
Progress of Polymers from Renewable Resources: Furans, Vegetable assessments of bio-based polymers. Polym. Degrad. Stab. 2013, 98,
Oils, and Polysaccharides. Chem. Rev. 2016, 116, 1637−1669. 1898−1907.
(25) Meier, M. A. R.; Metzger, J. O.; Schubert, U. S. Plant oil (49) Yates, M. R.; Barlow, C. Y. Life cycle assessments of
renewable resources as green alternatives in polymer science. Chem. biodegradable, commercial biopolymers-A critical review. Resour.
Soc. Rev. 2007, 36, 1788−1802. Conserv. Recy. 2013, 78, 54−66.
(26) Chen, G. Q.; Patel, M. K. Plastics Derived from Biological (50) Ellen MacArthur Foundation and McKinsey & Company. The
Sources: Present and Future: A Technical and Environmental Review. New Plastics Economy: Rethinking the Future of Plastics. https://
Chem. Rev. 2012, 112, 2082−2099. newplasticseconomy.org/ (accessed August 5, 2017).
(27) Williams, C. K.; Hillmyer, M. A. Polymers from renewable (51) Anastas, P. T.; Lankey, R. L. Life cycle assessment and green
resources: A perspective for a special issue of polymer reviews. Polym. chemistry: the yin and yang of industrial ecology. Green Chem. 2000, 2,
Rev. 2008, 48, 1−10. 289−295.
(28) Song, J. H.; Murphy, R. J.; Narayan, R.; Davies, G. B. H. (52) Anastas, P. W. J. Green Chemistry: Theory and Practice; Oxford
Biodegradable and compostable alternatives to conventional plastics. University Press, 2000.
Philos. Trans. R. Soc., B 2009, 364, 2127−2139. (53) Gonzalez, M. A.; Smith, R. L. A methodology to evaluate
(29) Gross, R. A.; Kalra, B. Biodegradable polymers for the process sustainability. Environ. Prog. 2003, 22, 269−276.
environment. Science 2002, 297, 803−807. (54) Guinee, J. Handbook on life cycle assessment - Operational
(30) Griffin, G. Chemistry and Technology of Biodegradable Polymers; guide to the ISO standards. Int. J. Life Cycle Assess. 2001, 6, 255−255.
Springer: Netherlands, 1994. (55) Ecoinvent Database: http://www.ecoinvent.org/ (accessed
(31) Swift, G. Directions for Environmentally Biodegradable Polymer December 22, 2016).
Research. Acc. Chem. Res. 1993, 26, 105−110. (56) Bare, J. TRACI 2.0: the tool for the reduction and assessment of
(32) Meadows, D. Leverage points. Places to Intervene in a System. chemical and other environmental impacts 2.0. Clean Technol. Environ.
http://drbalcom.pbworks.com/w/file/fetch/35173014/Leverage_ Policy 2011, 13, 687−696.
Points.pdf (accessed May 11, 2017). (57) Dornburg, V.; Lewandowski, I.; Patel, M. Comparing the Land
(33) Corma, A.; Iborra, S.; Velty, A. Chemical routes for the Requirements, Energy Savings, and Greenhouse Gas Emissions
transformation of biomass into chemicals. Chem. Rev. 2007, 107, Reduction of Biobased Polymers and Bioenergy. J. Ind. Ecol. 2003,
2411−2502. 7, 93−116.
(34) Rose, M.; Palkovits, R. Cellulose-Based Sustainable Polymers: (58) Kaenzig, J.; H, G.; Rocher, M. et al. Proceedings of the 2nd World
State of the Art and Future Trends. Macromol. Rapid Commun. 2011, Conference and Technology Exhibition on Biomass For Energy, Industry
32, 1299−1311. and Climate Change, Rome, Italy, 2004.
(35) Kumar, M. N. V. R.; Muzzarelli, R. A. A.; Muzzarelli, C.; (59) Tabone, M. D.; Cregg, J. J.; Beckman, E. J.; Landis, A. E.
Sashiwa, H.; Domb, A. J. Chitosan chemistry and pharmaceutical Sustainability Metrics: Life Cycle Assessment and Green Design in
perspectives. Chem. Rev. 2004, 104, 6017−6084. Polymers. Environ. Sci. Technol. 2010, 44, 8264−8269.
(36) Samir, M. A. S. A.; Alloin, F.; Dufresne, A. Review of recent (60) NatureWorks. http://www.natureworksllc.com/ (accessed
research into cellulosic whiskers, their properties and their application February 10, 2017).
in nanocomposite field. Biomacromolecules 2005, 6, 612−626. (61) Auras, R.; Harte, B.; Selke, S. An overview of polylactides as
(37) Rinaudo, M. Chitin and chitosan: Properties and applications. packaging materials. Macromol. Biosci. 2004, 4, 835−864.
Prog. Polym. Sci. 2006, 31, 603−632. (62) Showa Denko. http://www.showa-denko.com/news/bionolle-
(38) Menon, V.; Rao, M. Trends in bioconversion of lignocellulose: the-pioneer-in-biodegradable/ (accessed February 10, 2017).
Biofuels, platform chemicals & biorefinery concept. Prog. Energy (63) Avantium. https://www.avantium.com/yxy/ (accessed February
Combust. Sci. 2012, 38, 522−550. 10, 2017).
(39) Advancing Sustainable Materials Management 2014 Fact Sheet. (64) Braskem. http://www.braskem.com/site.aspx/Im-greenTM-
https://www.epa.gov/sites/production/files/2016-11/documents/ Polyethylene (accessed February 10, 2017).
2014_smmfactsheet_508.pdf (accessed May 11, 2017). (65) Mirel. http://www.mirelplastics.com/ (accessed February 10,
(40) Themelis, N. J.; Castaldi, M. J.; Bhatti, J.; Arsova, L., Energy and 2017).
Economics Value of Non-Recycled Plastics (NRP) and Municipal (66) Arkema. https://www.arkema.com/en/products/product-
Solid Wastes (MSW) that are Currently Landfilled in the Fifty States. finder/range-viewer/Rilsan-Polyamide-Family/ (accessed February
Columbia University, Earth Engineering Center, 2011. 10, 2017).
(41) Francis, R. Recycling of Polymers: Methods, Characterization and (67) DSM. https://www.dsm.com/products/ecopaxx/en_US/home.
Applications; Wiley-VCH, 2017. html (accessed February 10, 2017).
(42) Subramanian, P. M. Plastics recycling and waste management in (68) DuPont. http://sorona.com/ (accessed February 10, 2017).
the US. Resour. Conserv. Recy. 2000, 28, 253−263. (69) Novamont. http://www.novamont.com/eng/mater-bi (accessed
(43) Anastas, P. T.; Zimmerman, J. B. Peer Reviewed: Design February 10, 2017).
through the 12 principles of green engineering. Environ. Sci. Technol. (70) Coca-Cola. http://www.coca-colacompany.com/plantbottle-
2003, 37, 94A−101A. technology (accessed February 10, 2017).

873 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

(71) Empower Materials. http://www.empowermaterials.com/ (ac- (92) Cheng, K.-K.; Zhao, X.-B.; Zeng, J.; Zhang, J.-A. Biotechno-
cessed February 10, 2017). logical production of succinic acid: current state and perspectives.
(72) Saudi Aramco. http://www.novomer.com/novomer-announces- Biofuels, Bioprod. Biorefin. 2012, 6, 302−318.
commerc ial-la unch-converge%C2% AE -poly ols- broa d-use- (93) Beauprez, J. J.; De Mey, M.; Soetaert, W. K. Microbial succinic
polyurethanes-industry (accessed February 10, 2017). acid production: Natural versus metabolic engineered producers.
(73) Weiss, M.; Haufe, J.; Carus, M.; Brandao, M.; Bringezu, S.; Process Biochem. 2010, 45, 1103−1114.
Hermann, B.; Patel, M. K. A Review of the Environmental Impacts of (94) Xu, Q.; Li, S.; Huang, H.; Wen, J. Key technologies for the
Biobased Materials. J. Ind. Ecol. 2012, 16, S169−S181. industrial production of fumaric acid by fermentation. Biotechnol. Adv.
(74) Miller, S. A. Sustainable polymers: replacing polymers derived 2012, 30, 1685−1696.
from fossil fuels. Polym. Chem. 2014, 5, 3117−3118. (95) Okabe, M.; Lies, D.; Kanamasa, S.; Park, E. Y. Biotechnological
(75) Tsuji, H.; Suzuyoshi, K. Environmental degradation of production of itaconic acid and its biosynthesis in Aspergillus terreus.
biodegradable polyesters 1. Poly(ε-caprolactone), poly[(R)-3-hydrox- Appl. Microbiol. Biotechnol. 2009, 84, 597−606.
ybutyrate], and poly(L-lactide) films in controlled static seawater. (96) Bohnet, M. Ullmann’s Encyclopedia of Industrial Chemistry; John
Polym. Degrad. Stab. 2002, 75, 347−355. Wiley and Sons, Inc., 2003.
(76) Tsuji, H.; Suzuyoshi, K. Environmental degradation of (97) Deng, Y.; Ma, L.; Mao, Y. Biological production of adipic acid
biodegradable polyesters 2. Poly(ε-caprolactone), poly[(R)-3-hydrox- from renewable substrates: Current and future methods. Biochem. Eng.
ybutyrate], and poly(L-lactide) films in natural dynamic seawater. J. 2016, 105, 16−26.
Polym. Degrad. Stab. 2002, 75, 357−365. (98) Maisonneuve, L.; Lebarbé, T.; Grau, E.; Cramail, H. Structure−
(77) Fiorentino, G.; Ripa, M.; Ulgiati, S. Chemicals from biomass: properties relationship of fatty acid-based thermoplastics as synthetic
technological versus environmental feasibility. A review. Biofuels, polymer mimics. Polym. Chem. 2013, 4, 5472−5517.
Bioprod. Biorefin. 2017, 11, 195−214. (99) Goldbach, V.; Roesle, P.; Mecking, S. Catalytic Isomerizing ω-
(78) Cherubini, F.; Strømman, A. H. Chemicals from lignocellulosic Functionalization of Fatty Acids. ACS Catal. 2015, 5, 5951−5972.
biomass: opportunities, perspectives, and potential of biorefinery (100) Stempfle, F.; Quinzler, D.; Heckler, I.; Mecking, S. Long-Chain
systems. Biofuels, Bioprod. Biorefin. 2011, 5, 548−561. Linear C19and C23Monomers and Polycondensates from Unsaturated
(79) Delidovich, I.; Hausoul, P. J. C.; Deng, L.; Pfützenreuter, R.; Fatty Acid Esters. Macromolecules 2011, 44, 4159−4166.
Rose, M.; Palkovits, R. Alternative Monomers Based on Lignocellulose (101) Picataggio, S.; Rohrer, T.; Deanda, K.; Lanning, D.; Reynolds,
and Their Use for Polymer Production. Chem. Rev. 2016, 116, 1540− R.; Mielenz, J.; Eirich, L. D. Metabolic Engineering of Candida-
1599. Tropicalis for the Production of Long-Chain Dicarboxylic-Acids. Nat.
(80) Wyman, C. E. Potential Synergies and Challenges in Refining Biotechnol. 1992, 10, 894−898.
Cellulosic Biomass to Fuels, Chemicals, and Power. Biotechnol. Prog. (102) Stempfle, F.; Ortmann, P.; Mecking, S. Long-Chain Aliphatic
2003, 19, 254−262. Polymers To Bridge the Gap between Semicrystalline Polyolefins and
(81) Xia, Y.; Larock, R. C. Vegetable oil-based polymeric materials: Traditional Polycondensates. Chem. Rev. 2016, 116, 4597−4641.
synthesis, properties, and applications. Green Chem. 2010, 12, 1893− (103) de Jong, E.; Dam, M. A.; Sipos, L.; Gruter, G. J. M.
1909. Furandicarboxylic Acid (FDCA), A Versatile Building Block for a Very
(82) Foley, P. M.; Beach, E. S.; Zimmerman, J. B. Algae as a source of Interesting Class of Polyesters. ACS Symposium Series; American
renewable chemicals: opportunities and challenges. Green Chem. 2011, Chemical Society: Washington, DC, 2012; Vol. 1105, Chapter 1, pp
13, 1399−1405. 1−13.10.1021/bk-2012-1105.ch001
(83) Taherimehr, M.; Pescarmona, P. P. Green polycarbonates (104) Eerhart, A. J. J. E.; Faaij, A. P. C.; Patel, M. K. Replacing fossil
prepared by the copolymerization of CO2 with epoxides. J. Appl. based PET with biobased PEF; process analysis, energy and GHG
Polym. Sci. 2014, 131, 41141. balance. Energy Environ. Sci. 2012, 5, 6407−6422.
(84) Bhatia, S. K.; Bhatia, R. K.; Yang, Y.-H. Biosynthesis of (105) Papageorgiou, G. Z.; Papageorgiou, D. G.; Terzopoulou, Z.;
polyesters and polyamide building blocks using microbial fermentation Bikiaris, D. N. Production of bio-based 2,5-furan dicarboxylate
and biotransformation. Rev. Environ. Sci. Bio/Technol. 2016, 15, 639− polyesters: Recent progress and critical aspects in their synthesis and
663. thermal properties. Eur. Polym. J. 2016, 83, 202−229.
(85) Besson, M.; Gallezot, P.; Pinel, C. Conversion of Biomass into (106) van Putten, R. J.; van der Waal, J. C.; de Jong, E.; Rasrendra, C.
Chemicals over Metal Catalysts. Chem. Rev. 2014, 114, 1827−1870. B.; Heeres, H. J.; de Vries, J. G. Hydroxymethylfurfural, A Versatile
(86) Schwartz, T. J.; O’Neill, B. J.; Shanks, B. H.; Dumesic, J. A. Platform Chemical Made from Renewable Resources. Chem. Rev.
Bridging the Chemical and Biological Catalysis Gap: Challenges and 2013, 113, 1499−1597.
Outlooks for Producing Sustainable Chemicals. ACS Catal. 2014, 4, (107) Binder, J. B.; Raines, R. T. Simple Chemical Transformation of
2060−2069. Lignocellulosic Biomass into Furans for Fuels and Chemicals. J. Am.
(87) Llevot, A.; Dannecker, P.-K.; von Czapiewski, M.; Over, L. C.; Chem. Soc. 2009, 131, 1979−1985.
Söyler, Z.; Meier, M. A. R. Renewability is not Enough: Recent (108) da Costa Lopes, A. M.; Bogel-Lukasik, R. Acidic Ionic Liquids
Advances in the Sustainable Synthesis of Biomass-Derived Monomers as Sustainable Approach of Cellulose and Lignocellulosic Biomass
and Polymers. Chem. - Eur. J. 2016, 22, 11510−11521. Conversion without Additional Catalysts. ChemSusChem 2015, 8,
(88) Vilela, C.; Sousa, A. F.; Fonseca, A. C.; Serra, A. C.; Coelho, J. F. 947−965.
J.; Freire, C. S. R.; Silvestre, A. J. D. The quest for sustainable (109) Davis, S. E.; Houk, L. R.; Tamargo, E. C.; Datye, A. K.; Davis,
polyesters − insights into the future. Polym. Chem. 2014, 5, 3119− R. J. Oxidation of 5-hydroxymethylfurfural over supported Pt, Pd and
3141. Au catalysts. Catal. Today 2011, 160, 55−60.
(89) Gregory, G. L.; López-Vidal, E. M.; Buchard, A. Polymers from (110) Mishra, D. K.; Lee, H. J.; Kim, J.; Lee, H.-S.; Cho, J. K.; Suh,
sugars: cyclic monomer synthesis, ring-opening polymerisation, Y.-W.; Yi, Y.; Kim, Y. J. MnCo2O4 spinel supported ruthenium
material properties and applications. Chem. Commun. 2017, 53, catalyst for air-oxidation of HMF to FDCA under aqueous phase and
2198−2217. base-free conditions. Green Chem. 2017, 19, 1619−1623.
(90) Werpy, T.; Petersen, G. Top Value Added Chemicals from (111) Liu, B.; Zhang, Z. One-Pot Conversion of Carbohydrates into
Biomass. Vol. I - Results of Screening for Potential Candidates from Sugars Furan Derivatives via Furfural and 5-Hydroxylmethylfurfural as
and Synthesis Gas; U. S. Department of Energy, 2004. Intermediates. ChemSusChem 2016, 9, 2015−2036.
(91) Bozell, J. J.; Petersen, G. R. Technology development for the (112) Carlini, C.; Patrono, P.; Galletti, A. M. R.; Sbrana, G.; Zima, V.
production of biobased products from biorefinery carbohydratesthe Selective oxidation of 5-hydroxymethyl-2-furaldehyde to furan-2,5-
US Department of Energy’s “Top 10” revisited. Green Chem. 2010, 12, dicarboxaldehyde by catalytic systems based on vanadyl phosphate.
539−554. Appl. Catal., A 2005, 289, 197−204.

874 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

(113) Ribeiro, M. L.; Schuchardt, U. Cooperative effect of cobalt (136) Lee, C. S.; Aroua, M. K.; Daud, W. M. A. W.; Cognet, P.;
acetylacetonate and silica in the catalytic cyclization and oxidation of Peres-Lucchese, Y.; Fabre, P. L.; Reynes, O.; Latapie, L. A review:
fructose to 2,5-furandicarboxylic acid. Catal. Commun. 2003, 4, 83−86. Conversion of bioglycerol into 1,3-propanediol via biological and
(114) Sheldon, R. A. Green and sustainable manufacture of chemicals chemical method. Renewable Sustainable Energy Rev. 2015, 42, 963−
from biomass: state of the art. Green Chem. 2014, 16, 950−963. 972.
(115) Lange, J.-P.; van der Heide, E.; van Buijtenen, J.; Price, R. (137) DuPont Tate & Lyle BioProducts. http://www.
Furfural-A Promising Platform for Lignocellulosic Biofuels. Chem- duponttateandlyle.com/susterra (accessed May 2, 2017).
SusChem 2012, 5, 150−166. (138) Alonso-Fagúndez, N.; Granados, M. L.; Mariscal, R.; Ojeda, M.
(116) Banerjee, A.; Dick, G. R.; Yoshino, T.; Kanan, M. W. Carbon Selective Conversion of Furfural to Maleic Anhydride and Furan with
dioxide utilization via carbonate-promoted C−H carboxylation. Nature VOx/Al2O3Catalysts. ChemSusChem 2012, 5, 1984−1990.
2016, 531, 215−219. (139) Li, X.; Ho, B.; Zhang, Y. Selective aerobic oxidation of furfural
(117) Xie, G.; West, T. P. Citric acid production by Aspergillus niger to maleic anhydride with heterogeneous Mo−V−O catalysts. Green
ATCC 9142 from a treated ethanol fermentation co-product using Chem. 2016, 18, 2976−2980.
solid-state fermentation. Lett. Appl. Microbiol. 2009, 48, 639−644. (140) Lv, G.; Chen, C.; Lu, B.; Li, J.; Yang, Y.; Chen, C.; Deng, T.;
(118) Ilmen, M.; Koivuranta, K.; Ruohonen, L.; Suominen, P.; Zhu, Y.; Hou, X. Vanadium-oxo immobilized onto Schiff base modified
Penttila, M. Efficient Production of L-Lactic Acid from Xylose by graphene oxide for efficient catalytic oxidation of 5-hydroxymethyl-
Pichia stipitis. Appl. Environ. Microbiol. 2007, 73, 117−123. furfural and furfural into maleic anhydride. RSC Adv. 2016, 6,
(119) Zhang, D.; Hillmyer, M. A.; Tolman, W. B. A New Synthetic 101277−101282.
Route to Poly[3-hydroxypropionic acid] (P[3-HP]): Ring-Opening (141) Li, X.; Zhang, Y. The conversion of 5-hydroxymethyl furfural
Polymerization of 3-HP Macrocyclic Esters. Macromolecules 2004, 37, (HMF) to maleic anhydride with vanadium-based heterogeneous
8198−8200. catalysts. Green Chem. 2016, 18, 643−647.
(120) Ebata, H.; Toshima, K.; Matsumura, S. Lipase-catalyzed (142) bioamber. https://www.bio-amber.com/bioamber/en/
synthesis and curing of high-molecular-weight polyricinoleate. Macro- products (accessed May 2, 2017).
mol. Biosci. 2007, 7, 798−803. (143) Yim, H.; Haselbeck, R.; Niu, W.; Pujol-Baxley, C.; Burgard, A.;
(121) Slivniak, R.; Domb, A. J. Lactic acid and ricinoleic acid based Boldt, J.; Khandurina, J.; Trawick, J. D.; Osterhout, R. E.; Stephen, R.;
copolyesters. Macromolecules 2005, 38, 5545−5553. et al. Metabolic engineering of Escherichia coli for direct production of
(122) Shikanov, A.; Vaisman, B.; Shikanov, S.; Domb, A. J. Efficacy of 1,4-butanediol. Nat. Chem. Biol. 2011, 7, 445−452.
poly(sebacic acid-co-ricinoleic acid) biodegradable delivery system for (144) Burgard, A.; Burk, M. J.; Osterhout, R.; Van Dien, S.; Yim, H.
intratumoral delivery of paclitaxel. J. Biomed. Mater. Res., Part A 2010, Development of a commercial scale process for production of 1,4-
92, 1283−1291. butanediol from sugar. Curr. Opin. Biotechnol. 2016, 42, 118−125.
(123) Kassaian, J.-M. Ullmann’s Encyclopedia of Industrial Chemistry; (145) Biodiesel from Triglycerides via Transesterification; Springer:
VCH: Weinheim, Germany, 2002. London, 2008.
(124) Dhamaniya, S.; Jacob, J. Synthesis and characterization of (146) Zhang, J.; Li, J.-b.; Wu, S.-B.; Liu, Y. Advances in the Catalytic
polyesters based on tartaric acid derivatives. Polymer 2010, 51, 5392− Production and Utilization of Sorbitol. Ind. Eng. Chem. Res. 2013, 52,
5399. 11799−11815.
(125) Anderson, M. http://www.coca-colacompany.com/stories/ (147) Tathod, A.; Kane, T.; Sanil, E. S.; Dhepe, P. L. Solid base
great-things-come-in-innovative-packaging-an-introduction-to- supported metal catalysts for the oxidation and hydrogenation of
plantbottle-packaging (accessed May 2, 2017). sugars. J. Mol. Catal. A: Chem. 2014, 388−389, 90−99.
(126) Liu, X.; Wang, X.; Yao, S.; Jiang, Y.; Guan, J.; Mu, X. Recent (148) Dietrich, K.; Hernandez-Mejia, C.; Verschuren, P.;
advances in the production of polyols from lignocellulosic biomass and Rothenberg, G.; Shiju, N. R. One-Pot Selective Conversion of
biomass-derived compounds. RSC Adv. 2014, 4, 49501−49520. Hemicellulose to Xylitol. Org. Process Res. Dev. 2017, 21, 165−170.
(127) Yue, H.; Zhao, Y.; Ma, X.; Gong, J. Ethylene glycol: properties, (149) Akinterinwa, O.; Khankal, R.; Cirino, P. C. Metabolic
synthesis, and applications. Chem. Soc. Rev. 2012, 41, 4218−4244. engineering for bioproduction of sugar alcohols. Curr. Opin. Biotechnol.
(128) Darent, E.; Jager, W. W. Hydrogenolysis of glycerol. U.S. 2008, 19, 461−467.
Patent 6,080,898, 2000. (150) Nair, N. U.; Zhao, H. Evolution in Reverse: Engineering a D-
(129) Dasari, M. A.; Kiatsimkul, P. P.; Sutterlin, W. R.; Suppes, G. J. Xylose-Specific Xylose Reductase. ChemBioChem 2008, 9, 1213−1215.
Low-pressure hydrogenolysis of glycerol to propylene glycol. Appl. (151) Ladero, V.; Ramos, A.; Wiersma, A.; Goffin, P.; Schanck, A.;
Catal., A 2005, 281, 225−231. Kleerebezem, M.; Hugenholtz, J.; Smid, E. J.; Hols, P. High-Level
(130) Primo, A.; Concepción, P.; Corma, A. Synergy between the Production of the Low-Calorie Sugar Sorbitol by Lactobacillus
metal nanoparticles and the support for the hydrogenation of plantarum through Metabolic Engineering. Appl. Environ. Microbiol.
functionalized carboxylic acids to diols on Ru/TiO2. Chem. Commun. 2007, 73, 1864−1872.
2011, 47, 3613−3615. (152) Polaert, I.; Felix, M. C.; Fornasero, M.; Marcotte, S.; Buvat, J.
(131) Mao, B.-W.; Cai, Z.-Z.; Huang, M.-Y.; Jiang, Y.-Y. Hydro- C.; Estel, L. A greener process for isosorbide production: Kinetic study
genation of carboxylic acids catalyzed by magnesia-supported poly- of the catalytic dehydration of pure sorbitol under microwave. Chem.
?-aminopropylsiloxane-Ru complex. Polym. Adv. Technol. 2003, 14, Eng. J. 2013, 222, 228−239.
278−281. (153) de Almeida, R. M.; Li, J. R.; Nederlof, C.; O’Connor, P.;
(132) Takeda, Y.; Shoji, T.; Watanabe, H.; Tamura, M.; Nakagawa, Makkee, M.; Moulijn, J. A. Cellulose Conversion to Isosorbide in
Y.; Okumura, K.; Tomishige, K. Selective Hydrogenation of Lactic Molten Salt hydrate Media. ChemSusChem 2010, 3, 325−328.
Acid to 1,2-Propanediol over Highly Active Ruthenium-Molybdenum (154) Fenouillot, F.; Rousseau, A.; Colomines, G.; Saint-Loup, R.;
Oxide Catalysts. ChemSusChem 2015, 8, 1170−1178. Pascault, J. P. Polymers from renewable 1,4:3,6-dianhydrohexitols
(133) Niu, W.; Guo, J. Stereospecific Microbial Conversion of Lactic (isosorbide, isomannide and isoidide): A review. Prog. Polym. Sci.
Acid into 1,2-Propanediol. ACS Synth. Biol. 2015, 4, 378−382. 2010, 35, 578−622.
(134) Saxena, R. K.; Anand, P.; Saran, S.; Isar, J. Microbial (155) Dusselier, M.; Van Wouwe, P.; Dewaele, A.; Makshina, E.; Sels,
production of 1,3-propanediol: Recent developments and emerging B. F. Lactic acid as a platform chemical in the biobased economy: the
opportunities. Biotechnol. Adv. 2009, 27, 895−913. role of chemocatalysis. Energy Environ. Sci. 2013, 6, 1415−1442.
(135) Sivasankaran, C.; Govindaraj, K.; Mani, J. Bio-conversion of (156) Upare, P. P.; Hwang, Y. K.; Chang, J.-S.; Hwang, D. W.
glycerol into commercial production of 1,3-propanediol: A Review. J. Synthesis of Lactide from Alkyl Lactate via a Prepolymer Route. Ind.
Env. Biol. 2016, 37, 1539−1543. Eng. Chem. Res. 2012, 51, 4837−4842.

875 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

(157) Dusselier, M.; Van Wouwe, P.; Dewaele, A.; Jacobs, P. A.; Sels, (178) Jaffredo, C. G.; Guillaume, S. M. Benzyl β-malolactonate
B. F. Shape-selective zeolite catalysis for bioplastics production. Science polymers: a long story with recent advances. Polym. Chem. 2014, 5,
2015, 349, 78−80. 4168−4194.
(158) Dischert, W.; Colomb, C.; Soucaille, P. Fermentation Process (179) Duda, A.; Kowalski, A. In Handbook of Ring-Opening
for Producing Glycolic Acid. U.S. 20120178136A1, 2012. Polymerization; Dubois, P., Coulembier, O., Raquez, J., Eds.;
(159) Zhang, J.; Liu, X.; Sun, M.; Ma, X.; Han, Y. Direct Conversion WileyVCH: Weinheim, Germany, 2009; pp 1−51.
of Cellulose to Glycolic Acid with a Phosphomolybdic Acid Catalyst in (180) Strandman, S.; Gautrot, J. E.; Zhu, X. X. Recent advances in
a Water Medium. ACS Catal. 2012, 2, 1698−1702. entropy-driven ring-opening polymerizations. Polym. Chem. 2011, 2,
(160) Xiong, M. Y.; Schneiderman, D. K.; Bates, F. S.; Hillmyer, M. 791−799.
A.; Zhang, K. C. Scalable production of mechanically tunable block (181) Witt, T.; Mecking, S. Large-ring lactones from plant oils. Green
polymers from sugar. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 8357− Chem. 2013, 15, 2361−2364.
8362. (182) Pepels, M. P. F.; Koeken, R. A. C.; van der Linden, S. J. J.;
(161) Schneiderman, D. K.; Vanderlaan, M. E.; Mannion, A. M.; Heise, A.; Duchateau, R. Mimicking (Linear) Low-Density Poly-
Panthani, T. R.; Batiste, D. C.; Wang, J. Z.; Bates, F. S.; Macosko, C. ethylenes Using Modified Polymacrolactones. Macromolecules 2015,
W.; Hillmyer, M. A. Chemically Recyclable Biobased Polyurethanes. 48, 4779−4792.
ACS Macro Lett. 2016, 5, 515−518. (183) Shaikh, A.-A. G.; Sivaram, S. Organic Carbonates†. Chem. Rev.
(162) Brutman, J. P.; De Hoe, G. X.; Schneiderman, D. K.; Le, T. N.; 1996, 96, 951−976.
Hillmyer, M. A. Renewable, Degradable, and Chemically Recyclable (184) Sakakura, T.; Kohno, K. The synthesis of organic carbonates
Cross-Linked Elastomers. Ind. Eng. Chem. Res. 2016, 55, 11097− from carbon dioxide. Chem. Commun. 2009, 0, 1312−1330.
11106. (185) Choi, J.-C.; Kohno, K.; Ohshima, Y.; Yasuda, H.; Sakakura, T.
(163) Meng, Q.; Hou, M.; Liu, H.; Song, J.; Han, B. Synthesis of Tin- or titanium-catalyzed dimethyl carbonate synthesis from carbon
ketones from biomass-derived feedstock. Nat. Commun. 2017, 8, dioxide and methanol: Large promotion by a small amount of triflate
14190−14197. salts. Catal. Commun. 2008, 9, 1630−1633.
(164) Buntara, T.; Noel, S.; Phua, P. H.; Melian-Cabrera, I.; de Vries, (186) Rokicki, G. Aliphatic cyclic carbonates and spiroorthocar-
J. G.; Heeres, H. J. Caprolactam from Renewable Resources: Catalytic bonates as monomers. Prog. Polym. Sci. 2000, 25, 259−342.
Conversion of 5-Hydroxymethylfurfural into Caprolactone. Angew. (187) Burk, R. M.; Roof, M. B. A Safe and Efficient Method for
Chem., Int. Ed. 2011, 50, 7083−7087. Conversion of 1,2-Diols and 1,3-Diols to Cyclic Carbonates Utilizing
(165) Quilter, H. C.; Hutchby, M.; Davidson, M. G.; Jones, M. D. Triphosgene. Tetrahedron Lett. 1993, 34, 395−398.
Polymerisation of a terpene-derived lactone: a bio-based alternative to (188) Basel, Y.; Hassner, A. Di-tert-butyl dicarbonate and 4-
(dimethylamino)pyridine revisited. Their reactions with amines and
ε-caprolactone. Polym. Chem. 2017, 8, 833−837.
alcohols. J. Org. Chem. 2000, 65, 6368−6380.
(166) Zhang, D. H.; Hillmyer, M. A.; Tolman, W. B. Catalytic
(189) Mindemark, J.; Bowden, T. Synthesis and polymerization of
polymerization of a cyclic ester derived from a ″cool″ natural
alkyl halide-functional cyclic carbonates. Polymer 2011, 52, 5716−
precursor. Biomacromolecules 2005, 6, 2091−2095.
5722.
(167) Lowe, J. R.; Martello, M. T.; Tolman, W. B.; Hillmyer, M. A.
(190) Sanders, D. P.; Fukushima, K.; Coady, D. J.; Nelson, A.;
Functional biorenewable polyesters from carvone-derived lactones.
Fujiwara, M.; Yasumoto, M.; Hedrick, J. L. A Simple and Efficient
Polym. Chem. 2011, 2, 702−708.
Synthesis of Functionalized Cyclic Carbonate Monomers Using a
(168) Tang, M.; White, A. J. P.; Stevens, M. M.; Williams, C. K.
Versatile Pentafluorophenyl Ester Intermediate. J. Am. Chem. Soc.
Biomaterials from sugars: ring-opening polymerization of a carbohy- 2010, 132, 14724−14726.
drate lactone. Chem. Commun. 2009, 0, 941−943. (191) Mikami, K.; Lonnecker, A. T.; Gustafson, T. P.; Zinnel, N. F.;
(169) Romero Zaliz, C. L.; Varela, O. Facile synthesis of a d- Pai, P. J.; Russell, D. H.; Wooley, K. L. Polycarbonates Derived from
galactono-1,6-lactone derivative, a precursor of a copolyester. Glucose via an Organocatalytic Approach. J. Am. Chem. Soc. 2013, 135,
Carbohydr. Res. 2006, 341, 2973−2977. 6826−6829.
(170) Urakami, H.; Guan, Z. Living Ring-Opening Polymerization of (192) Ariga, T.; Takata, T.; Endo, T. Cationic ring-opening
a Carbohydrate-Derived Lactone for the Synthesis of Protein-Resistant polymerization of cyclic carbonates with alkyl halides to yield
Biomaterials. Biomacromolecules 2008, 9, 592−597. polycarbonate without the ether unit by suppression of elimination
(171) Bozell, J. J.; Petersen, G. R. Technology development for the of carbon dioxide. Macromolecules 1997, 30, 737−744.
production of biobased products from biorefinery carbohydrates-the (193) Chung, K.; Banik, S. M.; De Crisci, A. G.; Pearson, D. M.;
US Department of Energy’s ″Top 10″ revisited. Green Chem. 2010, 12, Blake, T. R.; Olsson, J. V.; Ingram, A. J.; Zare, R. N.; Waymouth, R. M.
539−554. Chemoselective Pd-Catalyzed Oxidation of Polyols: Synthetic Scope
(172) Shao, Z.; Li, C.; Di, X.; Xiao, Z.; Liang, C. Aqueous-Phase and Mechanistic Studies. J. Am. Chem. Soc. 2013, 135, 7593−7602.
Hydrogenation of Succinic Acid to γ-Butyrolactone and Tetrahy- (194) Painter, R. M.; Pearson, D. M.; Waymouth, R. M. Selective
drofuran over Pd/C, Re/C, and Pd−Re/C Catalysts. Ind. Eng. Chem. Catalytic Oxidation of Glycerol to Dihydroxyacetone. Angew. Chem.,
Res. 2014, 53, 9638−9645. Int. Ed. 2010, 49, 9456−9459.
(173) Budroni, G.; Corma, A. Gold and gold−platinum as active and (195) Simon, J.; Olsson, J. V.; Kim, H.; Tenney, I. F.; Waymouth, R.
selective catalyst for biomass conversion: Synthesis of γ-butyrolactone M. Semicrystalline Dihydroxyacetone Copolymers Derived from
and one-pot synthesis of pyrrolidone. J. Catal. 2008, 257, 403−408. Glycerol. Macromolecules 2012, 45, 9275−9281.
(174) Yokota, K.; Hirabayashi, T. J.P. Patent 04049288 A, (196) Pagliaro, M.; Ciriminna, R.; Kimura, H.; Rossi, M.; Della Pina,
Manufacture α-Methylene-γ-butyrolactone, 1992. C. From Glycerol to Value-Added Products. Angew. Chem., Int. Ed.
(175) Manzer, L. E. Catalytic synthesis of α-methylene-γ- 2007, 46, 4434−4440.
valerolactone: a biomass-derived acrylic monomer. Appl. Catal., A (197) Zelikin, A. N.; Zawaneh, P. N.; Putnam, D. A Functionalizable
2004, 272, 249−256. Biomaterial Based on Dihydroxyacetone, an Intermediate of Glucose
(176) Liu, D. J.; Chen, E. Y. X. Organocatalysis in biorefining for Metabolism. Biomacromolecules 2006, 7, 3239−3244.
biomass conversion and upgrading. Green Chem. 2014, 16, 964−981. (198) Wang, L. S.; Jiang, X. S.; Wang, H.; Cheng, S. X.; Zhuo, R. X.
(177) Tang, X.; Hong, M.; Falivene, L.; Caporaso, L.; Cavallo, L.; Preparation and cytotoxicity of novel aliphatic polycarbonate
Chen, E. Y. X. The Quest for Converting Biorenewable Bifunctional α- synthesized from dihydroxyacetone. Chin. Chem. Lett. 2005, 16,
Methylene-γ-butyrolactone into Degradable and Recyclable Polyester: 572−574.
Controlling Vinyl-Addition/Ring-Opening/Cross-Linking Pathways. J. (199) Brignou, P.; Priebe Gil, M.; Casagrande, O.; Carpentier, J.-F.
Am. Chem. Soc. 2016, 138, 14326−14337. o.; Guillaume, S. M. Polycarbonates Derived from Green Acids: Ring-

876 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

Opening Polymerization of Seven-Membered Cyclic Carbonates. (219) Kricheldorf, H. R.; Behnken, G.; Schwarz, G. Telechelic
Macromolecules 2010, 43, 8007−8017. polyesters of ethane diol and adipic or sebacic acid by means of
(200) Reithofer, M. R.; Sum, Y. N.; Zhang, Y. Synthesis of cyclic bismuth carboxylates as non-toxic catalysts. Polymer 2005, 46, 11219−
carbonates with carbon dioxide and cesium carbonate. Green Chem. 11224.
2013, 15, 2086−2090. (220) Foy, E.; Farrell, J. B.; Higginbotham, C. L. Synthesis of linear
(201) Bobbink, F. D.; Gruszka, W.; Hulla, M.; Das, S.; Dyson, P. J. aliphatic polycarbonate macroglycols using dimethylcarbonate. J. Appl.
Synthesis of cyclic carbonates from diols and CO2catalyzed by Polym. Sci. 2009, 111, 217−227.
carbenes. Chem. Commun. 2016, 52, 10787−10790. (221) Mutlu, H.; Ruiz, J.; Solleder, S. C.; Meier, M. A. R. TBD
(202) Gregory, G. L.; Ulmann, M.; Buchard, A. Synthesis of 6- catalysis with dimethyl carbonate: a fruitful and sustainable alliance.
membered cyclic carbonates from 1,3-diols and low CO2 pressure: a Green Chem. 2012, 14, 1728−1735.
novel mild strategy to replace phosgene reagents. RSC Adv. 2015, 5, (222) Sun, J.; Kuckling, D. Synthesis of high-molecular-weight
39404−39408. aliphatic polycarbonates by organo-catalysis. Polym. Chem. 2016, 7,
(203) Gregory, G. L.; Jenisch, L. M.; Charles, B.; Kociok-Kohn, G.; 1642−1649.
Buchard, A. Polymers from Sugars and CO2: Synthesis and (223) Zhu, J.; Cai, J.; Xie, W.; Chen, P.-H.; Gazzano, M.; Scandola,
Polymerization of a D-Mannose-Based Cyclic Carbonate. Macro- M.; Gross, R. A. Poly(butylene 2,5-furan dicarboxylate), a Biobased
molecules 2016, 49, 7165−7169. Alternative to PBT: Synthesis, Physical Properties, and Crystal
(204) Pinto, L. D.; Dupont, J.; de Souza, R. F.; Bernardo-Gusmão, K. Structure. Macromolecules 2013, 46, 796−804.
Catalytic asymmetric epoxidation of limonene using manganese Schiff- (224) Cao, A. Studies on syntheses and physical characterization of
base complexes immobilized in ionic liquids. Catal. Commun. 2008, 9, biodegradable aliphatic poly(butylene succinate-co-ε-caprolactone)s.
135−139. Polymer 2002, 43, 671−679.
(205) Winkler, M.; Romain, C.; Meier, M. A. R.; Williams, C. K. (225) Shirahama, H.; Kawaguchi, Y.; Aludin, M. S.; Yasuda, H.
Renewable polycarbonates and polyesters from 1,4-cyclohexadiene. Synthesis and enzymatic degradation of high molecular weight
Green Chem. 2015, 17, 300−306. aliphatic polyesters. J. Appl. Polym. Sci. 2001, 80, 340−347.
(206) Mutlu, H.; Hofsäß, R.; Montenegro, R. E.; Meier, M. A. R. Self- (226) Pokharkar, V.; Sivaram, S. Poly(Alkylene Carbonate)S by the
metathesis of fatty acid methyl esters: full conversion by choosing the Carbonate Interchange Reaction of Aliphatic Diols with Dimethyl
appropriate plant oil. RSC Adv. 2013, 3, 4927−4934. Carbonate - Synthesis and Characterization. Polymer 1995, 36, 4851−
(207) Biermann, U.; Sehlinger, A.; Meier, M. A. R.; Metzger, J. O. 4854.
Catalytic copolymerization of methyl 9,10-epoxystearate and cyclic (227) Wang, Z. Q.; Yang, X. G.; Liu, S. Y.; Hu, J.; Zhang, H.; Wang,
anhydrides under neat conditions. Eur. J. Lipid Sci. Technol. 2016, 118, G. Y. One-pot synthesis of high-molecular-weight aliphatic poly-
104−110. carbonates via melt transesterification of diphenyl carbonate and diols
(208) Bell, B. M.; Briggs, J. R.; Campbell, R. M.; Chambers, S. M.; using Zn(OAc)(2) as a catalyst. RSC Adv. 2015, 5, 87311−87319.
Gaarenstroom, P. D.; Hippler, J. G.; Hook, B. D.; Kearns, K.; Kenney, (228) Nakajima, T.; Tsukamoto, K.; Gyobu, S.; Yoshida, F.; Sato, M.;
J. M.; Kruper, W. J.; et al. Glycerin as a Renewable Feedstock for Watanabe, N.; Kageyama, K.; Kuwata, M.; Moriyama, N.; Matsumoto,
Epichlorohydrin Production. The GTE Process. Clean: Soil, Air, Water H.; et al. U.S. Patent 10/363,648, Polymerization catalyst for polyester,
2008, 36, 657−661. polyester produced with the same, and process for producing
(209) Zhou, C. H.; Zhao, H.; Tong, D. S.; Wu, L. M.; Yu, W. H. polyester, 2006.
Recent Advances in Catalytic Conversion of Glycerol. Catal. Rev.: Sci. (229) Masubuchi, T.; Ueno, E. U.S. Patent 2010/0292497,
Eng. 2013, 55, 369−453. Polycarbonate diol with ease of reaction stabilization, 2010.
(210) Hu, Y.; Qiao, L.; Qin, Y.; Zhao, X.; Chen, X.; Wang, X.; Wang, (230) Montaudo, G.; Rizzarelli, P. Synthesis and enzymatic
F. Synthesis and Stabilization of Novel Aliphatic Polycarbonate from degradation of aliphatic copolyesters. Polym. Degrad. Stab. 2000, 70,
Renewable Resource. Macromolecules 2009, 42, 9251−9254. 305−314.
(211) Longo, J. M.; Sanford, M. J.; Coates, G. W. Ring-Opening (231) Cao, A.; Okamura, T.; Nakayama, K.; Inoue, Y.; Masuda, T.
Copolymerization of Epoxides and Cyclic Anhydrides with Discrete Studies on syntheses and physical properties of biodegradable aliphatic
Metal Complexes: Structure−Property Relationships. Chem. Rev. poly(butylene succinate-co-ethylene succinate)s and poly(butylene
2016, 116, 15167−15197. succinate-co-diethylene glycol succinate)s. Polym. Degrad. Stab. 2002,
(212) Carothers, W. H. Studies on Polymerization and Ring 78, 107−117.
Formation. I. An Introduction to the General Theory of Condensation (232) Liu, C.; Liu, F.; Cai, J.; Xie, W.; Long, T. E.; Turner, S. R.;
Polymers. J. Am. Chem. Soc. 1929, 51, 2548−2559. Lyons, A.; Gross, R. A. Polymers from Fatty Acids: Poly(ω-hydroxyl
(213) Vilela, C.; Silvestre, A. J. D.; Meier, M. A. R. Plant Oil-Based tetradecanoic acid) Synthesis and Physico-Mechanical Studies.
Long-Chain C26Monomers and Their Polymers. Macromol. Chem. Biomacromolecules 2011, 12, 3291−3298.
Phys. 2012, 213, 2220−2227. (233) Winkler, M.; Lacerda, T. M.; Mack, F.; Meier, M. A. R.
(214) Noordover, B. A. J.; van Staalduinen, V. G.; Duchateau, R.; Renewable Polymers from Itaconic Acid by Polycondensation and
Koning, C. E.; Rolf, A. T. M.; Mak, M.; Heise, A.; Frissen, A. E.; van Ring-Opening-Metathesis Polymerization. Macromolecules 2015, 48,
Haveren, J. Co- and Terpolyesters Based on Isosorbide and Succinic 1398−1403.
Acid for Coating Applications: Synthesis and Characterization. (234) Kricheldorf, H. R. Syntheses of Biodegradable and
Biomacromolecules 2006, 7, 3406−3416. Biocompatible Polymers by Means of Bismuth Catalysts. Chem. Rev.
(215) Park, J. H.; Jeon, J. Y.; Lee, J. J.; Jang, Y.; Varghese, J. K.; Lee, 2009, 109, 5579−5594.
B. Y. Preparation of High-Molecular-Weight Aliphatic Polycarbonates (235) Buzin, P.; Lahcini, M.; Schellenberg, J.; Schellenberg, J.;
by Condensation Polymerization of Diols and Dimethyl Carbonate. Schwarz, G.; Kricheldorf, H. R. Aliphatic Polyesters by Bismuth
Macromolecules 2013, 46, 3301−3308. Triflate-Catalyzed Polycondensations of Dicarboxylic Acids and
(216) Zhang, J.; Zhu, W. X.; Li, C. C.; Zhang, D.; Xiao, Y. N.; Guan, Aliphatic Diols (vol 41, pg 8491, 2008). Macromolecules 2010, 43,
G. H.; Zheng, L. C. Effect of the biobased linear long-chain monomer 6511−6511.
on crystallization and biodegradation behaviors of poly(butylene (236) Buzin, P.; Lahcini, M.; Schwarz, G.; Kricheldorf, H. R. Aliphatic
carbonate)-based copolycarbonates. RSC Adv. 2015, 5, 2213−2222. Polyesters by Bismuth Triflate-Catalyzed Polycondensations of
(217) Tisserat, B.; O’Kuru, R. H.; Hwang, H.; Mohamed, A. A.; Dicarboxylic Acids and Aliphatic Diols. Macromolecules 2008, 41,
Holser, R. Glycerol citrate polyesters produced through heating 8491−8495.
without catalysis. J. Appl. Polym. Sci. 2012, 125, 3429−3437. (237) Vilela, C.; Silvestre, A. J. D.; Meier, M. A. R. Plant Oil-Based
(218) Odian, G. Principles of Polymerization, 4th ed; John Wiley & Long-Chain C-26 Monomers and Their Polymers. Macromol. Chem.
Sons: Hoboken, NJ, 2004. Phys. 2012, 213, 2220−2227.

877 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

(238) Sousa, A. F.; Gandini, A.; Silvestre, A. J. D.; Neto, C. P.; Cruz (258) Loos, K. Biocatalysis in Polymer Chemistry; Wiley-VCH:
Pinto, J. J. C.; Eckerman, C.; Holmbom, B. Novel Suberin-Based Weinheim, Germany, 2010.
Biopolyesters: From Synthesis to Properties. J. Polym. Sci., Part A: (259) Kobayashi, S. Enzymatic ring-opening polymerization and
Polym. Chem. 2011, 49, 2281−2291. polycondensation for the green synthesis of polyesters. Polym. Adv.
(239) Zhang, S.; Yang, J.; Liu, X.; Chang, J.; Cao, A. Synthesis and Technol. 2015, 26, 677−686.
Characterization of Poly(butylene succinate-co-butylene malate): A (260) Vouyiouka, S. N.; Topakas, E.; Katsini, A.; Papaspyrides, C. D.;
New Biodegradable Copolyester Bearing Hydroxyl Pendant Groups. Christakopoulos, P. A Green Route for the Preparation of Aliphatic
Biomacromolecules 2003, 4, 437−445. Polyesters via Lipase-catalyzed Prepolymerization and Low-temper-
(240) Hao, Q.; Yang, J.; Li, Q.; Li, Y.; Jia, L.; Fang, Q.; Cao, A. New ature Postpolymerization. Macromol. Mater. Eng. 2013, 298, 679−689.
Facile Approach to Novel Water-Soluble Aliphatic Poly(butylene (261) Kadokawa, J.; Habu, H.; Fukamachi, S.; Karasu, M.; Tagaya,
tartarate)s Bearing Reactive Hydroxyl Pendant Groups. Biomacromo- H.; Chiba, K. Direct polycondensation of carbon dioxide with xylylene
lecules 2005, 6, 3474−3480. glycols: a new method for the synthesis of polycarbonates. Macromol.
(241) Takasu, A.; Shibata, Y.; Narukawa, Y.; Hirabayashi, T. Rapid Commun. 1998, 19, 657−660.
Chemoselective dehydration polycondensations of dicarboxylic acids (262) Kadokawa, J.; Fukamachi, S.; Tagaya, H.; Chiba, K. Direct
and diols having pendant hydroxyl groups using the room temperature polycondensation of carbon dioxide with various diols using the
polycondensation technique. Macromolecules 2007, 40, 151−153. triphenylphosphine/bromotrichloromethane/N-cyclohexyl- N ′,N ′,N
(242) Hahn, C.; Wesselbaum, S.; Keul, H.; Möller, M. OH-functional ″,N ″-tetramethylguanidine system as condensing agent. Polym. J.
polyesters based on malic acid: Influence of the OH-groups onto the 2000, 32, 703−706.
thermal properties. Eur. Polym. J. 2013, 49, 217−227. (263) Tamura, M.; Ito, K.; Honda, M.; Nakagawa, Y.; Sugimoto, H.;
(243) Sokolsky-Papkov, M.; Langer, R.; Domb, A. J. Synthesis of Tomishige, K. Direct Copolymerization of CO2 and Diols. Sci. Rep.
aliphatic polyesters by polycondensation using inorganic acid as 2016, 6, 24038.
catalyst. Polym. Adv. Technol. 2011, 22, 502−511. (264) Oi, S.; Nemoto, K.; Matsuno, S.; Inoue, Y. Direct synthesis of
(244) Moyori, T.; Tang, T.; Takasu, A. Dehydration Polycondensa- polycarbonates from CO2, diols, and dihalides. Macromol. Rapid
tion of Dicarboxylic Acids and Diols Using Sublimating Strong Commun. 1994, 15, 133−137.
Brønsted Acids. Biomacromolecules 2012, 13, 1240−1243. (265) Chen, Z.; Hadjichristidis, N.; Feng, X.; Gnanou, Y. Cs2CO3-
(245) Schuchardt, U.; Sercheli, R.; Vargas, R. M. Transesterification promoted polycondensation of CO2with diols and dihalides for the
of vegetable oils: a review. J. Braz. Chem. Soc. 1998, 9, 199−210. synthesis of miscellaneous polycarbonates. Polym. Chem. 2016, 7,
(246) Türünç, O.; Meier, M. A. R. Fatty Acid Derived Monomers 4944−4952.
and Related Polymers Via Thiol-ene (Click) Additions. Macromol. (266) Bian, S.; Pagan, C.; Andrianova Artemyeva, A. A.; Du, G.
Rapid Commun. 2010, 31, 1822−1826. Andrianova “Artemyeva”, A. A.; Du, G. Synthesis of Polycarbonates
(247) Tang, D.; Noordover, B. A. J.; Sablong, R. J.; Koning, C. E. and Poly(ether carbonate)s Directly from Carbon Dioxide and Diols
Metal-free synthesis of novel biobased dihydroxyl-terminated aliphatic Promoted by a Cs2CO3/CH2Cl2 System. ACS Omega 2016, 1,
polyesters as building blocks for thermoplastic polyurethanes. J. Polym. 1049−1057.
Sci., Part A: Polym. Chem. 2011, 49, 2959−2968. (267) Guillaume, S. M.; Kirillov, E.; Sarazin, Y.; Carpentier, J. F.
(248) Sun, J. J.; Kuckling, D. Synthesis of high-molecular-weight Beyond Stereoselectivity, Switchable Catalysis: Some of the Last
aliphatic polycarbonates by organo-catalysis. Polym. Chem. 2016, 7, Frontier Challenges in Ring-Opening Polymerization of Cyclic Esters.
1642−1649. Chem. - Eur. J. 2015, 21, 7988−8003.
(249) Naik, P. U.; Refes, K.; Sadaka, F.; Brachais, C.-H.; Boni, G.; (268) Dechy-Cabaret, O.; Martin-Vaca, B.; Bourissou, D. Controlled
Couvercelle, J.-P.; Picquet, M.; Plasseraud, L. Organo-catalyzed ring-opening polymerization of lactide and glycolide. Chem. Rev. 2004,
synthesis of aliphatic polycarbonates in solvent-free conditions. 104, 6147−6176.
Polym. Chem. 2012, 3, 1475. (269) Albertsson, A. C.; Varma, I. K. Recent developments in ring
(250) Manabe, K.; Sun, X.-M.; Kobayashi, S. Dehydration Reactions opening polymerization of lactones for biomedical applications.
in Water. Surfactant-Type Brønsted Acid-Catalyzed Direct Esterifica- Biomacromolecules 2003, 4, 1466−1486.
tion of Carboxylic Acids with Alcohols in an Emulsion System. J. Am. (270) Jerome, C.; Lecomte, P. Recent advances in the synthesis of
Chem. Soc. 2001, 123, 10101−10102. aliphatic polyesters by ring-opening polymerization. Adv. Drug Delivery
(251) Manabe, K.; Iimura, S.; Sun, X.-M.; Kobayashi, S. Dehydration Rev. 2008, 60, 1056−1076.
Reactions in Water. Brønsted Acid−Surfactant-Combined Catalyst for (271) Gowda, R. R.; Chen, E. Y. X. Synthesis of β-methyl-α-
Ester, Ether, Thioether, and Dithioacetal Formation in Water. J. Am. methylene-γ-butyrolactone from biorenewable itaconic acid. Org.
Chem. Soc. 2002, 124, 11971−11978. Chem. Front. 2014, 1, 230−234.
(252) Takasu, A.; Takemoto, A.; Hirabayashi, T. Polycondensation of (272) van der Meulen, I.; Gubbels, E.; Huijser, S.; Sablong, R.;
dicarboxylic acids and diols in water catalyzed by surfactant-combined Koning, C. E.; Heise, A.; Duchateau, R. Catalytic Ring-Opening
catalysts and successive chain extension. Biomacromolecules 2006, 7, Polymerization of Renewable Macrolactones to High Molecular
6−9. Weight Polyethylene-like Polymers. Macromolecules 2011, 44, 4301−
(253) Sousa, A. F.; Silvestre, A. J. D.; Gandini, A.; Neto, C. P. 4305.
Synthesis of aliphatic suberin-like polyesters by ecofriendly catalytic (273) Duda, A., Kowalski, A. Thermodynamics and Kinetics of Ring-
systems. High Perform. Polym. 2012, 24, 4−8. Opening Polymerization. In Handbook of Ring-Opening Polymerization;
(254) Sousa, A. F.; Gandini, A.; Silvestre, A. J. D.; Pascoal Neto, C. Dubois, P., Coulembier, O., Raquez, J.-M., Eds.; Wiley-VCH:
Synthesis and Characterization of Novel Biopolyesters from Suberin Weinheim, Germany, 2009.
and Model Comonomers. ChemSusChem 2008, 1, 1020−1025. (274) Schneiderman, D. K.; Hillmyer, M. A. Aliphatic Polyester
(255) Pellis, A.; Herrero Acero, E.; Gardossi, L.; Ferrario, V.; Block Polymer Design. Macromolecules 2016, 49, 2419−2428.
Guebitz, G. M. Renewable building blocks for sustainable polyesters: (275) Olsén, P.; Odelius, K.; Albertsson, A.-C. Thermodynamic
new biotechnological routes for greener plastics. Polym. Int. 2016, 65, Presynthetic Considerations for Ring-Opening Polymerization. Bio-
861−871. macromolecules 2016, 17, 699−709.
(256) Gross, R. A.; Ganesh, M.; Lu, W. Enzyme-catalysis breathes (276) Duda, A.; Kowalski, A.; Libiszowski, J.; Penczek, S.
new life into polyester condensation polymerizations. Trends Thermodynamic and Kinetic Polymerizability of Cyclic Esters.
Biotechnol. 2010, 28, 435−443. Macromol. Symp. 2005, 224, 71−84.
(257) Yu, Y.; Wu, D.; Liu, C.; Zhao, Z.; Yang, Y.; Li, Q. Lipase/ (277) Houk, K. N.; Jabbari, A.; Hall, H. K.; Alemán, C. Why δ-
esterase-catalyzed synthesis of aliphatic polyesters via polycondensa- Valerolactone Polymerizes and γ-Butyrolactone Does Not. J. Org.
tion: A review. Process Biochem. 2012, 47, 1027−1036. Chem. 2008, 73, 2674−2678.

878 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

(278) Hong, M.; Chen, E. Y. X. Completely recyclable biopolymers Opening Polymerization of Trimethylene Carbonate. Chem. - Eur. J.
with linear and cyclic topologies via ring-opening polymerization of γ- 2008, 14, 8772−8775.
butyrolactone. Nat. Chem. 2015, 8, 42−49. (298) Brignou, P.; Carpentier, J.-F. o.; Guillaume, S. M. Metal- and
(279) van der Meulen, I.; Gubbels, E.; Huijser, S.; Sablong, R. l.; Organo-Catalyzed Ring-Opening Polymerization of α-Methyl-Tri-
Koning, C. E.; Heise, A.; Duchateau, R. Catalytic Ring-Opening methylene Carbonate: Insights into the Microstructure of the
Polymerization of Renewable Macrolactones to High Molecular Polycarbonate. Macromolecules 2011, 44, 5127−5135.
Weight Polyethylene-like Polymers. Macromolecules 2011, 44, 4301− (299) Stanford, M. J.; Dove, A. P. Stereocontrolled ring-opening
4305. polymerisation of lactide. Chem. Soc. Rev. 2010, 39, 486−494.
(280) Bouyahyi, M.; Duchateau, R. Metal-Based Catalysts for (300) Drumright, R. E.; Gruber, P. R.; Henton, D. E. Polylactic acid
Controlled Ring-Opening Polymerization of Macrolactones: High technology. Adv. Mater. 2000, 12, 1841−1846.
Molecular Weight and Well-Defined Copolymer Architectures. (301) Spassky, N.; Wisniewski, M.; Pluta, C.; Le Borgne, A. Highly
Macromolecules 2014, 47, 517−524. stereoelective polymerization of rac-(D,L)-lactide with a chiral schiff’s
(281) Bouyahyi, M.; Pepels, M. P. F.; Heise, A.; Duchateau, R. ω- base/aluminium alkoxide initiator. Macromol. Chem. Phys. 1996, 197,
Pentandecalactone Polymerization and ω-Pentadecalactone/ε-Capro- 2627−2637.
lactone Copolymerization Reactions Using Organic Catalysts. Macro- (302) Ovitt, T. M.; Coates, G. W. Stereoselective ring-opening
molecules 2012, 45, 3356−3366. polymerization of rac-lactide with a single-site, racemic aluminum
(282) Tempelaar, S.; Mespouille, L.; Coulembier, O.; Dubois, P.; alkoxide catalyst: Synthesis of stereoblock poly(lactic acid). J. Polym.
Dove, A. P. Synthesis and post-polymerisation modifications of Sci., Part A: Polym. Chem. 2000, 38, 4686−4692.
aliphatic poly(carbonate)s prepared by ring-opening polymerisation. (303) Ovitt, T. M.; Coates, G. W. Stereochemistry of lactide
Chem. Soc. Rev. 2013, 42, 1312−1336. polymerization with chiral catalysts: New opportunities for stereo-
(283) Venkataraman, S.; Ng, V. W. L.; Coady, D. J.; Horn, H. W.; control using polymer exchange mechanisms. J. Am. Chem. Soc. 2002,
Jones, G. O.; Fung, T. S.; Sardon, H.; Waymouth, R. M.; Hedrick, J. L.; 124, 1316−1326.
Yang, Y. Y. A Simple and Facile Approach to AliphaticN-Substituted (304) Zhong, Z. Y.; Dijkstra, P. J.; Feijen, J. [(salen)Al]-mediated,
Functional Eight-Membered Cyclic Carbonates and Their Organo- controlled and stereoselective ring-opening polymerization of lactide
catalytic Polymerization. J. Am. Chem. Soc. 2015, 137, 13851−13860. in solution and without solvent: Synthesis of highly isotactic
(284) Chang, Y. A.; Rudenko, A. E.; Waymouth, R. M. Zwitterionic polylactide stereocopolymers from racemic D,L-lactide. Angew.
Ring-Opening Polymerization of N-Substituted Eight-Membered Chem., Int. Ed. 2002, 41, 4510−4513.
Cyclic Carbonates to Generate Cyclic Poly(carbonate)s. ACS Macro (305) Zhong, Z. Y.; Dijkstra, P. J.; Feijen, J. Controlled and
Lett. 2016, 5, 1162−1166. stereoselective polymerization of lactide: Kinetics, selectivity, and
(285) Kricheldorf, H. R.; Kreisersaunders, I. Polylactones 0.19. microstructures. J. Am. Chem. Soc. 2003, 125, 11291−11298.
Anionic-Polymerization of L-Lactide in Solution. Makromol. Chem. (306) Nomura, N.; Ishii, R.; Akakura, M.; Aoi, K. Stereoselective
1990, 191, 1057−1066. ring-opening polymerization of racemic lactide using aluminum-achiral
(286) Jedlinski, Z.; Walach, W.; Kurcok, P.; Adamus, G. Polymer-
ligand complexes: Exploration of a chain-end control mechanism. J.
ization of Lactones 0.12. Polymerization of L-Dilactide and L,D-
Am. Chem. Soc. 2002, 124, 5938−5939.
Dilactide in the Presence of Potassium Methoxide. Makromol. Chem.
(307) Tang, Z. H.; Chen, X. S.; Pang, X.; Yang, Y. K.; Zhang, X. F.;
1991, 192, 2051−2057.
Jing, X. B. Stereoselective polymerization of rac-lactide using a
(287) Quirk, R. P.; Zhuo, Q.; Jang, S. H.; Lee, Y.; Lizarraga, G.
monoethylaluminum Schiff base complex. Biomacromolecules 2004, 5,
Principles of Anionic Polymerization: An Introduction. ACS Symp. Ser.
965−970.
1998, 696, 2−27.
(308) Tang, Z.; Chen, X.; Yang, Y.; Pang, X.; Sun, J.; Zhang, X.; Jing,
(288) Kricheldorf, H. R.; Dunsing, R.; Albet, A. S. i. Makromol. Chem.
X. Stereoselective polymerization ofrac-lactide with a bulky aluminum/
1987, 188, 2453−2466.
(289) Bourissou, D.; Martin-Vaca, B.; Dumitrescu, A.; Graullier, M.; Schiff base complex. J. Polym. Sci., Part A: Polym. Chem. 2004, 42,
Lacombe, F. Controlled cationic polymerization of lactide. Macro- 5974−5982.
molecules 2005, 38, 9993−9998. (309) Nomura, N.; Ishii, R.; Yamamoto, Y.; Kondo, T. Stereo-
(290) Sarazin, Y.; Carpentier, J. F. Discrete Cationic Complexes for selective Ring-Opening Polymerization of a Racemic Lactide by Using
Ring-Opening Polymerization Catalysis of Cyclic Esters and Epoxides. Achiral Salen− and Homosalen−Aluminum Complexes. Chem. - Eur. J.
Chem. Rev. 2015, 115, 3564−3614. 2007, 13, 4433−4451.
(291) Amgoune, A.; Thomas, C. M.; Carpentier, J.-F. Controlled (310) Bakewell, C.; Cao, T. P. A.; Long, N.; Le Goff, X. F.; Auffrant,
ring-opening polymerization of lactide by group 3 metal complexes. A.; Williams, C. K. Yttrium Phosphasalen Initiators for rac-Lactide
Pure Appl. Chem. 2007, 79, 2013−2030. Polymerization: Excellent Rates and High Iso-Selectivities. J. Am.
(292) Sauer, A.; Kapelski, A.; Fliedel, C.; Dagorne, S.; Kol, M.; Chem. Soc. 2012, 134, 20577−20580.
Okuda, J. Structurally well-defined group 4 metal complexes as (311) Bakewell, C.; White, A. J. P.; Long, N. J.; Williams, C. K. Metal-
initiators for the ring-opening polymerization of lactide monomers. Size Influence in Iso-Selective Lactide Polymerization. Angew. Chem.,
Dalton Trans. 2013, 42, 9007−9023. Int. Ed. 2014, 53, 9226−9230.
(293) Carpentier, J. F. Rare-Earth Complexes Supported by Tripodal (312) Jones, M. D.; Hancock, S. L.; McKeown, P.; Schafer, P. M.;
Tetradentate Bis(phenolate) Ligands: A Privileged Class of Catalysts Buchard, A.; Thomas, L. H.; Mahon, M. F.; Lowe, J. P. Zirconium
for Ring-Opening Polymerization of Cyclic Esters. Organometallics complexes of bipyrrolidine derived salan ligands for the isoselective
2015, 34, 4175−4189. polymerisation of rac-lactide. Chem. Commun. 2014, 50, 15967−
(294) Dijkstra, P. J.; Du, H. Z.; Feijen, J. Single site catalysts for 15970.
stereoselective ring-opening polymerization of lactides. Polym. Chem. (313) Wang, H. B.; Ma, H. Y. Highly diastereoselective synthesis of
2011, 2, 520−527. chiral aminophenolate zinc complexes and isoselective polymerization
(295) Labet, M.; Thielemans, W. Synthesis of polycaprolactone: a of rac-lactide. Chem. Commun. 2013, 49, 8686−8688.
review. Chem. Soc. Rev. 2009, 38, 3484−3504. (314) Wang, H. B.; Yang, Y.; Ma, H. Y. Exploring Steric Effects in
(296) Ajellal, N.; Carpentier, J. F.; Guillaume, C.; Guillaume, S. M.; Diastereoselective Synthesis of Chiral Aminophenolate Zinc Com-
Helou, M.; Poirier, V.; Sarazin, Y.; Trifonov, A. Metal-catalyzed plexes and Stereoselective Ring-Opening Polymerization of rac-
immortal ring-opening polymerization of lactones, lactides and cyclic Lactide. Inorg. Chem. 2016, 55, 7356−7372.
carbonates. Dalton Trans. 2010, 39, 8363−8376. (315) Abbina, S.; Du, G. D. Zinc-Catalyzed Highly Isoselective Ring
(297) Helou, M.; Miserque, O.; Brusson, J.-M.; Carpentier, J.-F.; Opening Polymerization of rac-Lactide. ACS Macro Lett. 2014, 3,
Guillaume, S. M. Ultraproductive, Zinc-Mediated, Immortal Ring- 689−692.

879 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

(316) Mou, Z.; Liu, B.; Wang, M.; Xie, H.; Li, P.; Li, L.; Li, S.; Cui, D. (337) Spink, S. S.; Kazakov, O. I.; Kiesewetter, E. T.; Kiesewetter, M.
Isoselective ring-opening polymerization of rac-lactide initiated by K. Rate Accelerated Organocatalytic Ring-Opening Polymerization of
achiral heteroscorpionate zwitterionic zinc complexes. Chem. Commun. L-Lactide via the Application of a Bis(thiourea) H-bond Donating
2014, 50, 11411−11414. Cocatalyst. Macromolecules 2015, 48, 6127−6131.
(317) Lenz, R. W.; Marchessault, R. H. Bacterial Polyesters: (338) Fastnacht, K. V.; Spink, S. S.; Dharmaratne, N. U.; Pothupitiya,
Biosynthesis, Biodegradable Plastics and Biotechnology. Biomacromo- J. U.; Datta, P. P.; Kiesewetter, E. T.; Kiesewetter, M. K. Bis- and Tris-
lecules 2005, 6, 1−8. Urea H-Bond Donors for Ring-Opening Polymerization: Unprece-
(318) Carpentier, J. F. Discrete Metal Catalysts for Stereoselective dented Activity and Control from an Organocatalyst. ACS Macro Lett.
Ring-Opening Polymerization of Chiral Racemic beta-Lactones. 2016, 5, 982−986.
Macromol. Rapid Commun. 2010, 31, 1696−1705. (339) Makiguchi, K.; Yamanaka, T.; Kakuchi, T.; Terada, M.; Satoh,
(319) Wang, X.; Thevenon, A.; Brosmer, J. L.; Yu, I.; Khan, S. I.; T. Binaphthol-derived phosphoric acids as efficient chiral organo-
Mehrkhodavandi, P.; Diaconescu, P. L. Redox Control of Group 4 catalysts for the enantiomer-selective polymerization of rac-lactide.
Metal Ring-Opening Polymerization Activity towardl-Lactide and ε- Chem. Commun. 2014, 50, 2883−2885.
Caprolactone. J. Am. Chem. Soc. 2014, 136, 11264−11267. (340) Miyake, G. M.; Chen, E. Y. X. Cinchona Alkaloids as
(320) Biernesser, A. B.; Delle Chiaie, K. R.; Curley, J. B.; Byers, J. A. Stereoselective Organocatalysts for the Partial Kinetic Resolution
Block Copolymerization of Lactide and an Epoxide Facilitated by a Polymerization of rac-Lactide. Macromolecules 2011, 44, 4116−4124.
Redox Switchable Iron-Based Catalyst. Angew. Chem., Int. Ed. 2016, 55, (341) Zhu, J. B.; Chen, E. Y. X. From meso-Lactide to Isotactic
5251−5254. Polylactide: Epimerization by B/N Lewis Pairs and Kinetic Resolution
(321) Quan, S. M.; Wang, X. K.; Zhang, R. J.; Diaconescu, P. L. by Organic Catalysts. J. Am. Chem. Soc. 2015, 137, 12506−12509.
Redox Switchable Copolymerization of Cyclic Esters and Epoxides by (342) Sanchez-Sanchez, A.; Rivilla, I.; Agirre, M.; Basterretxea, A.;
a Zirconium Complex. Macromolecules 2016, 49, 6768−6778. Etxeberria, A.; Veloso, A.; Sardon, H.; Mecerreyes, D.; Cossío, F. P.
(322) Romain, C.; Williams, C. K. Chemoselective Polymerization Enantioselective Ring-Opening Polymerization of rac-Lactide Dictated
Control: From Mixed-Monomer Feedstock to Copolymers. Angew. by Densely Substituted Amino Acids. J. Am. Chem. Soc. 2017, 139,
Chem., Int. Ed. 2014, 53, 1607−1610. 4805−4814.
(323) Romain, C.; Zhu, Y.; Dingwall, P.; Paul, S.; Rzepa, H. S.; (343) Paul, S.; Zhu, Y.; Romain, C.; Brooks, R.; Saini, P. K.; Williams,
Buchard, A.; Williams, C. K. Chemoselective Polymerizations from C. K. Ring-opening copolymerization (ROCOP): synthesis and
Mixtures of Epoxide, Lactone, Anhydride, and Carbon Dioxide. J. Am. properties of polyesters and polycarbonates. Chem. Commun. 2015,
Chem. Soc. 2016, 138, 4120−4131. 51, 6459−6479.
(324) Nederberg, F.; Connor, E. F.; Möller, M.; Glauser, T.; Hedrick, (344) Coates, G. W.; Moore, D. R. Discrete metal-based catalysts for
J. L. New paradigms for organic catalysts: The first organocatalytic the copolymerization CO2 and epoxides: Discovery, reactivity,
living polymerization. Angew. Chem., Int. Ed. 2001, 40, 2712−2715. optimization, and mechanism. Angew. Chem., Int. Ed. 2004, 43,
(325) Kamber, N. E.; Jeong, W.; Waymouth, R. M.; Pratt, R. C.; 6618−6639.
Lohmeijer, B. G. G.; Hedrick, J. L. Organocatalytic Ring-Opening (345) Darensbourg, D. J. Making Plastics from Carbon Dioxide:
Polymerization. Chem. Rev. 2007, 107, 5813−5840. Salen Metal Complexes as Catalysts for the Production of
(326) Kiesewetter, M. K.; Shin, E. J.; Hedrick, J. L.; Waymouth, R. M. Polycarbonates from Epoxides and CO2. Chem. Rev. 2007, 107,
Organocatalysis: Opportunities and Challenges for Polymer Synthesis. 2388−2410.
Macromolecules 2010, 43, 2093−2107. (346) Klaus, S.; Lehenmeier, M. W.; Anderson, C. E.; Rieger, B.
(327) Brown, H. A.; Waymouth, R. M. Zwitterionic Ring-Opening Recent advances in CO2/epoxide copolymerizationNew strategies
Polymerization for the Synthesis of High Molecular Weight Cyclic and cooperative mechanisms. Coord. Chem. Rev. 2011, 255, 1460−
Polymers. Acc. Chem. Res. 2013, 46, 2585−2596. 1479.
(328) Dove, A. P. Metal-Free Catalysis in Ring-Opening Polymer- (347) Darensbourg, D. J.; Yeung, A. D. A concise review of
ization. In Handbook of Ring-Opening Polymerization; Dubois, P., computational studies of the carbon dioxide-epoxide copolymerization
Coulembier, O., Raquez, J.-M., Eds., Wiley-VCH: Weinheim, 2009; reactions. Polym. Chem. 2014, 5, 3949−3962.
Chapter 14. (348) Darensbourg, D. J.; Wilson, S. J. What’s new with CO2?
(329) Thomas, C.; Bibal, B. Hydrogen-bonding organocatalysts for Recent advances in its copolymerization with oxiranes. Green Chem.
ring-opening polymerization. Green Chem. 2014, 16, 1687−1699. 2012, 14, 2665−2671.
(330) Ottou, W. N.; Sardon, H.; Mecerreyes, D.; Vignolle, J.; Taton, (349) Lu, X. B.; Ren, W. M.; Wu, G. P. CO2 Copolymers from
D. Update and challenges in organo-mediated polymerization Epoxides: Catalyst Activity, Product Selectivity, and Stereochemistry
reactions. Prog. Polym. Sci. 2016, 56, 64−115. Control. Acc. Chem. Res. 2012, 45, 1721−1735.
(331) Dove, A. P. Organic Catalysis for Ring-Opening Polymer- (350) Childers, M. I.; Longo, J. M.; Van Zee, N. J.; LaPointe, A. M.;
ization. ACS Macro Lett. 2012, 1, 1409−1412. Coates, G. W. Stereoselective Epoxide Polymerization and Copoly-
(332) Suriano, F.; Coulembier, O.; Hedrick, J. L.; Dubois, P. merization. Chem. Rev. 2014, 114, 8129−8152.
Functionalized cyclic carbonates: from synthesis and metal-free (351) Kember, M. R.; Buchard, A.; Williams, C. K. Catalysts for
catalyzed ring-opening polymerization to applications. Polym. Chem. CO2/epoxide copolymerisation. Chem. Commun. 2011, 47, 141−163.
2011, 2, 528−533. (352) Trott, G.; Saini, P. K.; Williams, C. K. Catalysts for CO2/
(333) Guillerm, B.; Lemaur, V.; Cornil, J.; Lazzaroni, R.; Dubois, P.; epoxide ring-opening copolymerization. Philos. Trans. R. Soc., A 2016,
Coulembier, O. Ammonium betaines: efficient ionic nucleophilic 374, 20150085.
catalysts for the ring-opening polymerization ofl-lactide and cyclic (353) Ree, M.; Bae, J. Y.; Jung, J. H.; Shin, T. J. A new
carbonates. Chem. Commun. 2014, 50, 10098−10101. copolymerization process leading to poly(propylene carbonate) with
(334) Naumann, S.; Thomas, A. W.; Dove, A. P. Highly Polarized a highly enhanced yield from carbon dioxide and propylene oxide. J.
Alkenes as Organocatalysts for the Polymerization of Lactones and Polym. Sci., Part A: Polym. Chem. 1999, 37, 1863−1876.
Trimethylene Carbonate. ACS Macro Lett. 2016, 5, 134−138. (354) Liu, Y.; Xiao, M.; Wang, S.; Xia, L.; Hang, D.; Cui, G.; Meng,
(335) Wang, Q.; Zhao, W.; He, J.; Zhang, Y.; Chen, E. Y. X. Living Y. Mechanism studies of terpolymerization of phthalic anhydride,
Ring-Opening Polymerization of Lactones byN-Heterocyclic Olefin/ propylene epoxide, and carbon dioxide catalyzed by ZnGA. RSC Adv.
Al(C6F5)3Lewis Pairs: Structures of Intermediates, Kinetics, and 2014, 4, 9503−9508.
Mechanism. Macromolecules 2017, 50, 123−136. (355) Chen, S.; Hua, Z.; Fang, Z.; Qi, G. Copolymerization of carbon
(336) Lin, B.; Waymouth, R. M. Urea Anions: Simple, Fast, and dioxide and propylene oxide with highly effective zinc
Selective Catalysts for Ring-Opening Polymerizations. J. Am. Chem. hexacyanocobaltate(III)-based coordination catalyst. Polymer 2004,
Soc. 2017, 139, 1645−1652. 45, 6519−6524.

880 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

(356) Sun, X.-K.; Zhang, X.-H.; Chen, S.; Du, B.-Y.; Wang, Q.; Fan, (376) Poirier, Y.; Nawrath, C.; Somerville, C. Production of
Z.-Q.; Qi, G.-R. One-pot terpolymerization of CO2, cyclohexene oxide polyhydroxyalkanoates, a family of biodegradable plastics and
and maleic anhydride using a highly active heterogeneous double elastomers, in bacteria and plants. Nat. Biotechnol. 1995, 13, 142−150.
metal cyanide complex catalyst. Polymer 2010, 51, 5719−5725. (377) Haywood, G. W.; Anderson, A. J.; Dawes, E. A. The
(357) Nakano, K.; Nakamura, M.; Nozaki, K. Alternating Importance of Phb-Synthase Substrate-Specificity in Polyhydroxyalka-
Copolymerization of Cyclohexene Oxide with Carbon Dioxide noate Synthesis by Alcaligenes-Eutrophus. FEMS Microbiol. Lett. 1989,
Catalyzed by (salalen)CrCl Complexes. Macromolecules 2009, 42, 57, 1−6.
6972−6980. (378) Asenjo, J. A.; Suk, J. S. Microbial Conversion of Methane into
(358) van Meerendonk, W. J.; Duchateau, R.; Koning, C. E.; Gruter, poly-β-hydroxybutyrate (PHB): Growth and intracellular product
G.-J. M. Unexpected Side Reactions and Chain Transfer for Zinc- accumulation in a type II methanotroph. J. Ferment. Technol. 1986, 64,
Catalyzed Copolymerization of Cyclohexene Oxide and Carbon 271−278.
Dioxide. Macromolecules 2005, 38, 7306−7313. (379) Panchal, B.; Bagdadi, A.; Roy, I. Polyhydroxyalkanoates: The
(359) Cohen, C. T.; Chu, T.; Coates, G. W. Cobalt Catalysts for the Natural Polymers Produced by Bacterial Fermentation. In Advances in
Alternating Copolymerization of Propylene Oxide and Carbon Natural Polymers: Composites and Nanocomposites; Springer: Berlin,
Dioxide: Combining High Activity and Selectivity. J. Am. Chem. Soc. 2013.
2005, 127, 10869−10878. (380) Singh Saharan, B.; Grewal, A.; Kumar, P. Biotechnological
(360) Cyriac, A.; Lee, S. H.; Varghese, J. K.; Park, E. S.; Park, J. H.; Production of Polyhydroxyalkanoates: A Review on Trends and Latest
Lee, B. Y. Immortal CO2/Propylene Oxide Copolymerization: Precise Developments. Chin. J. Bio. 2014, 2014, 1−18.
Control of Molecular Weight and Architecture of Various Block (381) Flanagan, J. C. A.; Myung, J.; Criddle, C. S.; Waymouth, R. M.
Copolymers. Macromolecules 2010, 43, 7398−7401. Poly(hydroxyalkanoate)s from Waste Biomass: A Combined Chem-
(361) Pescarmona, P. P.; Taherimehr, M. Challenges in the catalytic ical-Biological Approach. ChemistrySelect 2016, 1, 2327−2331.
synthesis of cyclic and polymeric carbonates from epoxides and CO2. (382) Myung, J.; Flanagan, J. C. A.; Waymouth, R. M.; Criddle, C. S.
Catal. Sci. Technol. 2012, 2, 2169−2187. Methane or methanol-oxidation dependent synthesis of poly(3-
(362) Nakano, K.; Kamada, T.; Nozaki, K. Selective Formation of hydroxybutyrate-co-3-hydroxyvalerate) by obligate type II methano-
Polycarbonate over Cyclic Carbonate: Copolymerization of Epoxides trophs. Process Biochem. 2016, 51, 561−567.
with Carbon Dioxide Catalyzed by a Cobalt(III) Complex with a (383) The use of methane waste as a feedstock for PHAs has been
Piperidinium End-Capping Arm. Angew. Chem., Int. Ed. 2006, 45, tested by Mango Materials. https://www.biobasedworldnews.com/
7274−7277. ask-the-industry-dr.-molly-morse-ceo-and-co-founder-mango-materials
(363) Noh, E. K.; Na, S. J.; S, S.; Kim, S.-W.; Lee, B. Y. Two (accessed May 8, 2017).
Components in a Molecule: Highly Efficient and Thermally Robust (384) Bugnicourt, E.; Cinelli, P.; Lazzeri, A.; Alvarez, V. The Main
Catalytic System for CO2/Epoxide Copolymerization. J. Am. Chem. Characteristics, Properties, Improvements, and Market Data of
Soc. 2007, 129, 8082−8083. Polyhydroxyalkanoates. In Handbook of Sustainable Polymers; Taylor
(364) S, S.; Min, J. K.; Seong, J. E.; Na, S. J.; Lee, B. Y. A Highly & Francis Group, 2015; Chapter 24.
Active and Recyclable Catalytic System for CO2/Propylene Oxide (385) Steinbüchel, A.; Valentin, H. E. Diversity of bacterial
Copolymerization. Angew. Chem., Int. Ed. 2008, 47, 7306−7309. polyhydroxyalkanoic acids. FEMS Microbiol. Lett. 1995, 128, 219−228.
(365) Byrne, C. M.; Allen, S. D.; Lobkovsky, E. B.; Coates, G. W. (386) Clark, J. H.; Farmer, T. J.; Herrero-Davila, L.; Sherwood, J.
Alternating Copolymerization of Limonene Oxide and Carbon Circular economy design considerations for research and process
Dioxide. J. Am. Chem. Soc. 2004, 126, 11404−11405. development in the chemical sciences. Green Chem. 2016, 18, 3914−
(366) Auriemma, F.; De Rosa, C.; Di Caprio, M. R.; Di Girolamo, R.; 3934.
Ellis, W. C.; Coates, G. W. Stereocomplexed Poly(Limonene (387) The term “Eco-efficiency” was coined by the World Business
Carbonate): A Unique Example of the Cocrystallization of Amorphous Council for Sustainable Development (WBCSD) in its 1992
Enantiomeric Polymers. Angew. Chem., Int. Ed. 2015, 54, 1215−1218. publication “Changing Course”.
(367) Jeske, R. C.; Rowley, J. M.; Coates, G. W. Pre-Rate- (388) Hopewell, J.; Dvorak, R.; Kosior, E. Plastics recycling:
Determining Selectivity in the Terpolymerization of Epoxides, Cyclic challenges and opportunities. Philos. Trans. R. Soc., B 2009, 364,
Anhydrides, and CO2: A One-Step Route to Diblock Copolymers. 2115−2126.
Angew. Chem., Int. Ed. 2008, 47, 6041−6044. (389) Shen, L.; Worrell, E.; Patel, M. K. Open-loop recycling: A LCA
(368) Saini, P. K.; Romain, C.; Zhu, Y.; Williams, C. K. Di- case study of PET bottle-to-fibre recycling. Resour. Conserv. Recy. 2010,
magnesium and zinc catalysts for the copolymerization of phthalic 55, 34−52.
anhydride and cyclohexene oxide. Polym. Chem. 2014, 5, 6068−6075. (390) Ohya, Y.; Takahashi, A.; Nagahama, K. Biodegradable
(369) Rehm, B. H. A. Bacterial polymers: biosynthesis, modifications Polymeric Assemblies for Biomedical Materials. Adv. Polym. Sci.
and applications. Nat. Rev. Microbiol. 2010, 8, 578−592. 2011, 247, 65−114.
(370) Chen, G. G.-Q. Plastics from Bacteria; Springer-Verlag: Berlin, (391) Pandey, S. K.; Haldar, C.; Patel, D. K.; Maiti, P. Biodegradable
2010. Polymers for Potential Delivery Systems for Therapeutics. Adv. Polym.
(371) Hazer, D. B.; Kılıçay, E.; Hazer, B. Poly(3-hydroxyalkanoate)s: Sci. 2013, 254, 169−202.
Diversification and biomedical applications. Mater. Sci. Eng., C 2012, (392) Brannigan, R. P.; Dove, A. P. Synthesis, properties and
32, 637−647. biomedical applications of hydrolytically degradable materials based on
(372) Li, Z.; Loh, X. J. Water soluble polyhydroxyalkanoates: future aliphatic polyesters and polycarbonates. Biomater. Sci. 2017, 5, 9−21.
materials for therapeutic applications. Chem. Soc. Rev. 2015, 44, 2865− (393) NAPCOR Recycling Rate Reports. http://www.napcor.com/
2879. PET/pet_reports.html (accessed May 8, 2017).
(373) Chen, G.-Q. A microbial polyhydroxyalkanoates (PHA) based (394) Data from Noone A., Collected PET bottles, Proceedings of
bio- and materials industry. Chem. Soc. Rev. 2009, 38, 2434−2446. 13th International Polyester Recycling Forum, 2008.
(374) Verlinden, R. A. J.; Hill, D. J.; Kenward, M. A.; Williams, C. D.; (395) Ignatyev, I. A.; Thielemans, W.; Vander Beke, B. Recycling of
Radecka, I. Bacterial synthesis of biodegradable polyhydroxyalka- Polymers: A Review. ChemSusChem 2014, 7, 1579−1593.
noates. J. Appl. Microbiol. 2007, 102, 1437−1449. (396) Fisher, M. M. Plastics Recycling; John Wiley & Sons, Inc:
(375) Myung, J.; Strong, N. I.; Galega, W. M.; Sundstrom, E. R.; Germany, 2003.
Flanagan, J. C. A.; Woo, S.-G.; Waymouth, R. M.; Criddle, C. S. (397) Nishida, H. Development of materials and technologies for
Disassembly and reassembly of polyhydroxyalkanoates: Recycling control of polymer recycling. Polym. J. 2011, 43, 435−447.
through abiotic depolymerization and biotic repolymerization. (398) loop Industries: Leading the sustainable plastic revolution.
Bioresour. Technol. 2014, 170, 167−174. http://www.loopindustries.com/en/ (accessed August 5, 2017).

881 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

(399) CARBIOS: Reinvent Polymers Lifecycle. https://carbios.fr/ (422) Lyu, S. P.; Schley, J.; Loy, B.; Lind, D.; Hobot, C.; Sparer, R.;
en/ (accessed August 5, 2017). Untereker, D. Kinetics and time-temperature equivalence of polymer
(400) Ioniqa Technologies. http://www.ioniqa.com/ (accessed degradation. Biomacromolecules 2007, 8, 2301−2310.
August 5, 2017). (423) Li, S. M.; Garreau, H.; Vert, M. Structure Property
(401) gren: Reduce Reuse Recycle. http://gr3n-recycling.com/ Relationships in the Case of the Degradation of Massive Aliphatic
(accessed August 5, 2017). Poly-(Alpha-Hydroxy Acids) in Aqueous-Media 0.1. Poly(Dl-Lactic
(402) Achilias, D. Material Recycling: Trends and Perspectives; InTech, Acid). J. Mater. Sci.: Mater. Med. 1990, 1, 123−130.
2012. (424) Codari, F.; Lazzari, S.; Soos, M.; Storti, G.; Morbidelli, M.;
(403) Paszun, D.; Spychaj, T. Chemical Recycling of Poly(ethylene Moscatelli, D. Kinetics of the hydrolytic degradation of poly(lactic
terephthalate). Ind. Eng. Chem. Res. 1997, 36, 1373−1383. acid). Polym. Degrad. Stab. 2012, 97, 2460−2466.
(404) Liu, F.-S.; Li, Z.; Yu, S.-T.; Cui, X.; Xie, C.-X.; Ge, X.-P. (425) Jung, J. H.; Ree, M.; Kim, H. Acid- and base-catalyzed
Methanolysis and Hydrolysis of Polycarbonate Under Moderate hydrolyses of aliphatic polycarbonates and polyesters. Catal. Today
Conditions. J. Polym. Environ. 2009, 17, 208−211. 2006, 115, 283−287.
(405) Achilias, D. S.; Redhwi, H. H.; Siddiqui, M. N.; Nikolaidis, A. (426) Shih, C. Chain-end scission in acid catalyzed hydrolysis of poly
K.; Bikiaris, D. N.; Karayannidis, G. P. Glycolytic depolymerization of (d,l-lactide) in solution. J. Controlled Release 1995, 34, 9−15.
PET waste in a microwave reactor. J. Appl. Polym. Sci. 2010, 118, (427) Batycky, R. P.; Hanes, J.; Langer, R.; Edwards, D. A. A
3066−3073. Theoretical Model of Erosion and Macromolecular Drug Release from
(406) Tsintzou, G. P.; Achilias, D. S. Chemical Recycling of Biodegrading Microspheres. J. Pharm. Sci. 1997, 86, 1464−1477.
Polycarbonate Based Wastes Using Alkaline Hydrolysis Under (428) van Nostrum, C. F.; Veldhuis, T. F. J.; Bos, G. W.; Hennink,
Microwave Irradiation. Waste Biomass Valorization 2013, 4, 3−7. W. E. Hydrolytic degradation of oligo(lactic acid): a kinetic and
(407) Shukla, S. R.; Harad, A. M.; Jawale, L. S. Recycling of waste mechanistic study. Polymer 2004, 45, 6779−6787.
PET into useful textile auxiliaries. Waste Manage. 2008, 28, 51−56. (429) Pereira, C. S. M.; Silva, V. M. T. M.; Rodrigues, A. E. Ethyl
(408) Horn, H. W.; Jones, G. O.; Wei, D. D. S.; Fukushima, K.; lactate as a solvent: Properties, applications and production processes
Lecuyer, J. M.; Coady, D. J.; Hedrick, J. L.; Rice, J. E. Mechanisms of − a review. Green Chem. 2011, 13, 2658−2671.
Organocatalytic Amidation and Trans-Esterification of Aromatic Esters (430) Bakker, H. R.; Kranz, M. T. C. Paint Composition Based on a
As a Model for the Depolymerization of Poly(ethylene) Terephthalate. Chemical Cross Linking System and/or Oxidative Drying, With
J. Phys. Chem. A 2012, 116, 12389−12398. Lactates As Solvent and Thinner. E.P. Patent 1023405 A1, 1999.
(409) Allen, R. D.; Bajjuri, K. M.; Hedrick, J. L.; Breyta, G.; Larson, (431) Morgan, J. D. Low-Temperature Lubricants. U.S. Patent
C. E. Methods and Materials for Depolymerizing Polyesters. U.S. 2389924 A, 1945.
Patent 9255194 B2, 2016. (432) Muse, J., Jr.; Colvin, H. A. Use of Ethyl Lactate As an Excipient
(410) Genta, M., Yano, F., Kondo, Y., Matsubara, W., Oomoto, S.
for Pharmaceutical Compositions. U.S. Patent 20050287179, 2005.
Technical Review; Mitsubishi Heavy Industries, Ltd., 2003, Vol. 40, pp
(433) Song, X.; Zhang, X.; Wang, H.; Liu, F.; Yu, S.; Liu, S.
1−4.
Methanolysis of poly(lactic acid) (PLA) catalyzed by ionic liquids.
(411) Jie, H.; Ke, H.; Qing, Z.; Lei, C.; Yongqiang, W.; Zibin, Z.
Polym. Degrad. Stab. 2013, 98, 2760−2764.
Study on depolymerization of polycarbonate in supercritical ethanol.
(434) Liu, H.; Song, X.; Liu, F.; Liu, S.; Yu, S. Ferric chloride as an
Polym. Degrad. Stab. 2006, 91, 2307−2314.
efficient and reusable catalyst for methanolysis of poly(lactic acid)
(412) Noritake, A.; Hori, M.; Shigematsu, M.; Tanahashi, M.
waste. J. Polym. Res. 2015, 22, 135.
Recycling of Polyethylene Terephthalate Using High-pressure Steam
(435) Filachione, E. M.; Lengel, J. H.; Fisher, C. H. Preparation of
Treatment. Polym. J. 2008, 40, 498−502.
(413) Watanabe, M.; Matsuo, Y.; Matsushita, T.; Inomata, H.; Methyl Lactate. Ind. Eng. Chem. 1945, 37, 388−390.
(436) Fliedel, C.; Vila-Viçosa, D.; Calhorda, M. J.; Dagorne, S.;
Miyake, T.; Hironaka, K. Chemical recycling of polycarbonate in high
pressure high temperature steam at 573 K. Polym. Degrad. Stab. 2009, Avilés, T. Dinuclear Zinc-N-Heterocyclic Carbene Complexes for
94, 2157−2162. Either the Controlled Ring-Opening Polymerization of Lactide or the
(414) Gioia, C.; Vannini, M.; Marchese, P.; Minesso, A.; Cavalieri, R.; Controlled Degradation of Polylactide Under Mild Conditions.
Colonna, M.; Celli, A. Sustainable polyesters for powder coating ChemCatChem 2014, 6, 1357−1367.
applications from recycled PET, isosorbide and succinic acid. Green (437) Brake, L. D. Recovery of polyhydroxy acids. U.S. Patent
Chem. 2014, 16, 1807−1815. 5264614 A, 1993.
(415) Gioia, C.; Minesso, A.; Cavalieri, R.; Marchese, P.; Celli, A.; (438) Brake, L. D. Preparation of alkyl esters by depolymerization of
Colonna, M. Powder coatings for indoor applications from renewable poly(hydroxy acids). U.S. Patent 5264617 A, 1993.
resources and recycled polymers. J. Coating. Technol. Res. 2015, 12, (439) Leibfarth, F. A.; Moreno, N.; Hawker, A. P.; Shand, J. D.
555−562. Transforming polylactide into value-added materials. J. Polym. Sci., Part
(416) Gioia, C.; Vannini, M.; Celli, A.; Colonna, M.; Minesso, A. A: Polym. Chem. 2012, 50, 4814−4822.
Chemical recycling of post-consumer compact discs towards novel (440) Petrus, R.; Bykowski, D.; Sobota, P. Solvothermal Alcoholysis
polymers for powder coating applications. RSC Adv. 2016, 6, 31462− Routes for Recycling Polylactide Waste as Lactic Acid Esters. ACS
31469. Catal. 2016, 6, 5222−5235.
(417) Jones, G. O.; Yuen, A.; Wojtecki, R. J.; Hedrick, J. L.; Garcia, J. (441) Delgado, P.; Sanz, M. T.; Beltrán, S.; Núñez, L. A. Ethyl lactate
M. Computational and experimental investigations of one-step production via esterification of lactic acid with ethanol combined with
conversion of poly(carbonate)s into value-added poly(aryl ether pervaporation. Chem. Eng. J. 2010, 165, 693−700.
sulfone)s. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 7722−7726. (442) Martino-Gauchi, G.; Teissier, R. Continuous Esterification
(418) Kumar, M. N. V. R. Handbook of Polyester Drug Delivery Process for the Preparation of Ethyl Lactate from Lactic Acid and
Systems; CRC Press, Taylor & Francis Group, 2016. Ethanol. U.S. Patent 20060041165 A1, 2004.
(419) Grayson, A. C. R.; Cima, M. J.; Langer, R. Size and (443) Plichta, A.; Lisowska, P.; Kundys, A.; Zychewicz, A.; Dębowski,
temperature effects on poly(lactic-co-glycolic acid) degradation and M.; Florjańczyk, Z. Chemical recycling of poly(lactic acid) via
microreservoir device performance. Biomaterials 2005, 26, 2137−2145. controlled degradation with protic (macro)molecules. Polym. Degrad.
(420) Tamada, J. A.; Langer, R. Erosion Kinetics of Hydrolytically Stab. 2014, 108, 288−296.
Degradable Polymers. Proc. Natl. Acad. Sci. U. S. A. 1993, 90, 552−556. (444) McNeill, I. C.; Leiper, H. A. Degradation studies of some
(421) Lazzari, S.; Codari, R.; Storti, G.; Morbidelli, M.; Moscatelli, D. polyesters and polycarbonates1. Polylactide: General features of the
Modeling the pH-dependent PLA oligomer degradation kinetics. degradation under programmed heating conditions. Polym. Degrad.
Polym. Degrad. Stab. 2014, 110, 80−90. Stab. 1985, 11, 267−285.

882 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

(445) McNeill, I. C.; Leiper, H. A. Degradation studies of some crotonic acid by catalytic thermal degradation. Polym. Degrad. Stab.
polyesters and polycarbonates2. Polylactide: Degradation under 2010, 95, 1375−1381.
isothermal conditions, thermal degradation mechanism and photolysis (467) Kim, K. J.; Doi, Y.; Abe, H. Effects of residual metal
of the polymer. Polym. Degrad. Stab. 1985, 11, 309−326. compounds and chain-end structure on thermal degradation of poly(3-
(446) Kopinke, F. D.; Remmler, M.; Mackenzie, K.; Moder, M.; hydroxybutyric acid). Polym. Degrad. Stab. 2006, 91, 769−777.
Wachsen, O. Thermal decomposition of biodegradable polyesters 0.2. (468) Merrick, J. M.; Doudoroff, M. Enzymatic Synthesis of Poly-β-
Poly(lactic acid). Polym. Degrad. Stab. 1996, 53, 329−342. Hydroxybutyric Acid in Bacteria. Nature 1961, 189, 890−892.
(447) Wachsen, O.; Platkowski, K.; Reichert, K. H. Thermal (469) Merrick, J. M.; Doudoroff, M. Depolymerization of Poly-Beta-
degradation of poly-l-lactidestudies on kinetics, modelling and Hydroxybutyrate by Intracellular Enzyme System. J. Bacteriol. 1964,
melt stabilisation. Polym. Degrad. Stab. 1997, 57, 87−94. 88, 60−71.
(448) Nishida, H.; Mori, T.; Hoshihara, S.; Fan, Y. J.; Shirai, Y.; (470) Eubeler, J. P.; Bernhard, M.; Knepper, T. P. Environmental
Endo, T. Effect of tin on poly(L-lactic acid) pyrolysis. Polym. Degrad. biodegradation of synthetic polymers II. Biodegradation of different
Stab. 2003, 81, 515−523. polymer groups. TrAC, Trends Anal. Chem. 2010, 29, 84−100.
(449) Mori, T.; Nishida, H.; Shirai, Y.; Endo, T. Effects of chain end (471) Eubeler, J. P.; Zok, S.; Bernhard, M.; Knepper, T. P.
structures on pyrolysis of poly(L-lactic acid) containing tin atoms. Environmental biodegradation of synthetic polymers I. Test method-
Polym. Degrad. Stab. 2004, 84, 243−251. ologies and procedures. TrAC, Trends Anal. Chem. 2009, 28, 1057−
(450) Noda, M.; Okuyama, H. Thermal catalytic depolymerization of 1072.
poly(L-Lactic acid) oligomer into LL-lactide: Effects of Al, Ti, Zn and (472) Tokiwa, Y.; Calabia, B. P.; Ugwu, C. U.; Aiba, S.
Zr compounds as catalysts. Chem. Pharm. Bull. 1999, 47, 467−471. Biodegradability of Plastics. Int. J. Mol. Sci. 2009, 10, 3722−3742.
(451) Fan, Y.; Nishida, H.; Mori, T.; Shirai, Y.; Endo, T. Thermal (473) Shah, A. A.; Hasan, F.; Hameed, A.; Ahmed, S. Biological
degradation of poly(l-lactide): effect of alkali earth metal oxides for degradation of plastics: A comprehensive review. Biotechnol. Adv. 2008,
selective l,l-lactide formation. Polymer 2004, 45, 1197−1205. 26, 246−265.
(452) Nishida, H.; Fan, Y. J.; Mori, T.; Oyagi, N.; Shirai, Y.; Endo, T. (474) Lucas, N.; Bienaime, C.; Belloy, C.; Queneudec, M.; Silvestre,
Feedstock recycling of flame-resisting poly(lactic acid)/aluminum F.; Nava-Saucedo, J.-E. Polymer biodegradation: Mechanisms and
hydroxide composite to L,L-lactide. Ind. Eng. Chem. Res. 2005, 44, estimation techniques − A review. Chemosphere 2008, 73, 429−442.
1433−1437. (475) Krzan, A.; Hemjinda, S.; Miertus, S.; Corti, A.; Chiellini, E.
(453) Motoyama, T.; Tsukegi, T.; Shirai, Y.; Nishida, H.; Endo, T. Standardization and certification in the area of environmentally
Effects of MgO catalyst on depolymerization of poly-L-lactic acid to degradable plastics. Polym. Degrad. Stab. 2006, 91, 2819−2833.
L,L-lactide. Polym. Degrad. Stab. 2007, 92, 1350−1358. (476) Moon, J.; Kim, M. Y.; Kim, B. M.; Lee, J. C.; Choi, M.-C.; Kim,
(454) Hocker, H.; Keul, H. Ring-Opening Polymerization and J. R. Estimation of the Microbial Degradation of Biodegradable
Depolymerization in Respective Polymers. Macromol. Symp. 1995, 98, Polymer, Poly(lactic acid) (PLA) with a Specific Gas Production Rate.
825−834. Macromol. Res. 2016, 24, 415−421.
(455) Nelissen, M.; Keul, H.; Hocker, H. Ring-Closing Depolyme- (477) Yagi, H.; Ninomiya, F.; Funabashi, M.; Kunioka, M. Anaerobic
rization of Poly(Epsilon-Caprolactone). Macromol. Chem. Phys. 1995, biodegradation tests of poly(lactic acid) and polycaprolactone using
196, 1645−1661. new evaluation system for methane fermentation in anaerobic sludge.
(456) Spanagel, E. W.; Carothers, W. H. Preparation of macrocyclic Polym. Degrad. Stab. 2009, 94, 1397−1404.
Lactones by depolymerization. J. Am. Chem. Soc. 1936, 58, 654−656. (478) Yagi, H.; Ninomiya, F.; Funabashi, M.; Kunioka, M.
(457) Spanagel, E. W.; Carothers, W. H. Macrocyclic Esters1. J. Am. Thermophilic anaerobic biodegradation test and analysis of eubacteria
Chem. Soc. 1935, 57, 929−934. involved in anaerobic biodegradation of four specified biodegradable
(458) Wood, B. R.; Semlyen, J. A.; Hodge, P. Preparation and polyesters. Polym. Degrad. Stab. 2013, 98, 1182−1187.
characterization of large ether-ester rings. Polymer 1997, 38, 2287− (479) Itävaara, M.; Karjomaa, S.; Selin, J. F. Biodegradation of
2290. polylactide in aerobic and anaerobic thermophilic conditions. Chemo-
(459) Ruddick, C. L.; Hodge, P.; Zhuo, Y.; Beddoes, R. L.; Helliwell, sphere 2002, 46, 879−885.
M. Cyclo-depolymerisation of polyundecanoate and related polyesters: (480) Massardier-Nageotte, V.; Pestre, C.; Cruard-Pradet, T.; Bayard,
characterisation of cyclic oligoundecanoates and related cyclic R. Aerobic and anaerobic biodegradability of polymer films and
oligoesters. J. Mater. Chem. 1999, 9, 2399−2405. physico-chemical characterization. Polym. Degrad. Stab. 2006, 91, 620−
(460) Hong, M.; Chen, E. Y. X. Towards Truly Sustainable Polymers: 627.
A Metal-Free Recyclable Polyester from Biorenewable Non-Strained γ- (481) Kale, G.; Auras, R.; Singh, S. P.; Narayan, R. Biodegradability
Butyrolactone. Angew. Chem., Int. Ed. 2016, 55, 4188−4193. of polylactide bottles in real and simulated composting conditions.
(461) Xiong, M.; Schneiderman, D. K.; Bates, F. S.; Hillmyer, M. A.; Polym. Test. 2007, 26, 1049−1061.
Zhang, K. Scalable production of mechanically tunable block polymers (482) Starnecker, A.; Menner, M. Assessment of biodegradability of
from sugar. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 8357−8362. plastics under simulated composting conditions in a laboratory test
(462) MacDonald, J. P.; Shaver, M. P. An aromatic/aliphatic system. Int. Biodeterior. Biodegrad. 1996, 37, 85−92.
polyester prepared via ring-opening polymerisation and its remarkably (483) González Petit, M.; Correa, Z.; Sabino, M. A. Degradation of a
selective and cyclable depolymerisation to monomer. Polym. Chem. Polycaprolactone/Eggshell Biocomposite in a Bioreactor. J. Polym.
2016, 7, 553−559. Environ. 2015, 23, 11−20.
(463) Akmal, D. Biodegradation of microbial polyesters P(3HB) and (484) Wing Hong, L.; Yu, J. Environmental factors and kinetics of
P(3HB-co-3HV) under the tropical climate environment. Polym. microbial degradation of poly(3-hydroxybutyrate-co-3-hydroxyvaler-
Degrad. Stab. 2003, 80, 513−518. ate) in an aqueous medium. J. Appl. Polym. Sci. 2003, 87, 205−213.
(464) Rostkowski, K. H.; Criddle, C. S.; Lepech, M. D. Cradle-to- (485) Sekiguchi, T.; Saika, A.; Nomura, K.; Watanabe, T.; Watanabe,
Gate Life Cycle Assessment for a Cradle-to-Cradle Cycle: Biogas-to- T.; Fujimoto, Y.; Enoki, M.; Sato, T.; Kato, C.; Kanehiro, H.
Bioplastic (and Back). Environ. Sci. Technol. 2012, 46, 9822−9829. Biodegradation of aliphatic polyesters soaked in deep seawaters and
(465) Grassie, N.; Murray, E. J.; Holmes, P. A. The thermal isolation of poly(ε-caprolactone)-degrading bacteria. Polym. Degrad.
degradation of poly(-(d)-β-hydroxybutyric acid): Part 1Identifica- Stab. 2011, 96, 1397−1403.
tion and quantitative analysis of products. Polym. Degrad. Stab. 1984, 6, (486) Hanphakphoom, S.; Maneewong, N.; Sukkhum, S.; Tokuyama,
47−61. S.; Kitpreechavanich, V. Characterization of poly(L-lactide)-degrading
(466) Ariffin, H.; Nishida, H.; Shirai, Y.; Hassan, M. A. Highly enzyme produced by thermophilic filamentous bacteria Laceyella
selective transformation of poly[(R)-3-hydroxybutyric acid] into trans- sacchari LP175. J. Gen. Appl. Microbiol. 2014, 60, 13−22.

883 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

(487) Sukkhum, S.; Tokuyama, S.; Kitpreechavanich, V. Poly(L- (506) He, F.; Li, S.; Vert, M.; Zhuo, R. Enzyme-catalyzed
Lactide)-Degrading Enzyme Production by Actinomadura keratinily- polymerization and degradation of copolymers prepared from ϵ-
tica T16−1 in 3 L Airlift Bioreactor and Its Degradation Ability for caprolactone and poly(ethylene glycol). Polymer 2003, 44, 5145−
Biological Recycle. J. Microbiol. Biotechnol. 2012, 22, 92−99. 5151.
(488) Fukushima, K.; Abbate, C.; Tabuani, D.; Gennari, M.; Camino, (507) He, F.; Li, S.; Garreau, H.; Vert, M.; Zhuo, R. Enzyme-
G. Biodegradation of poly(lactic acid) and its nanocomposites. Polym. catalyzed polymerization and degradation of copolyesters of ε-
Degrad. Stab. 2009, 94, 1646−1655. caprolactone and γ-butyrolactone. Polymer 2005, 46, 12682−12688.
(489) Lee, S. Y.; Lee, Y.; Wang, F. L. Chiral compounds from (508) Hwang, Y.; Ree, M.; Kim, H. Enzymatic degradation of
bacterial polyesters: Sugars to plastics to fine chemicals. Biotechnol. poly(propylene carbonate) and poly(propylene carbonate-co-ε-
Bioeng. 1999, 65, 363−368. caprolactone) synthesized via CO2 fixation. Catal. Today 2006, 115,
(490) Ren, Q.; Grubelnik, A.; Hoerler, M.; Ruth, K.; Hartmann, R.; 288−294.
Felber, H.; Zinn, M. Bacterial Poly(hydroxyalkanoates) as a Source of (509) Dai, L.; Xiao, S.; Shen, Y.; Qinshu, B.; He, J. The Synthesis of
Chiral Hydroxyalkanoic Acids. Biomacromolecules 2005, 6, 2290−2298. Cellulose-graft-poly (L-lactide) by Ring-opening Polymerization and
(491) Jendrossek, D. Microbial degradation of polyesters: a review on the Study of Its Degradability. Bull. Korean Chem. Soc. 2012, 33,
extracellular poly(hydroxyalkanoic acid) depolymerases. Polym. De- 4122−4126.
grad. Stab. 1998, 59, 317−325. (510) Rodríguez-Contreras, A.; Calafell-Monfort, M.; Marqués-
(492) Briese, B. H.; Jendrossek, D. Biological basis of enzyme- Calvo, M. S. Enzymatic degradation of poly(3-hydroxybutyrate-co-4-
catalyzed polyester degradation: 59 C-terminal amino acids of poly(3- hydroxybutyrate) by commercial lipases. Polym. Degrad. Stab. 2012,
hydroxybutyrate) (PHB) depolymerase a from pseudomonas 97, 597−604.
lemoignei are sufficient for PHB binding. Macromol. Symp. 1998, (511) Tsutsumi, C.; Hayase, N.; Nakagawa, K.; Tanaka, S.; Miyahara,
130, 205−216. Y. The enzymatic degradation of commercial biodegradable polymers
(493) Kasuya, K.; Ohura, T.; Masuda, K.; Doi, Y. Substrate and by some lipases and chemical degradation of them. Macromol. Symp.
binding specificities of bacterial polyhydroxybutyrate depolymerases. 2003, 197, 431−442.
Int. J. Biol. Macromol. 1999, 24, 329−336. (512) Vyner, M. C.; Li, A.; Amsden, B. G. The effect of
(494) Hiraishi, T.; Komiya, N.; Maeda, M. Y443F mutation in the poly(trimethylene carbonate) molecular weight on macrophage
substrate-binding domain of extracellular PHB depolymerase enhances behavior and enzyme adsorption and conformation. Biomaterials
its PHB adsorption and disruption abilities. Polym. Degrad. Stab. 2010, 2014, 35, 9041−9048.
95, 1370−1374. (513) Sommerfeld, S. D.; Zhang, Z.; Costache, M. C.; Vega, S. L.;
(495) Hiraishi, T.; Komiya, N.; Matsumoto, N.; Abe, H.; Fujita, M.; Kohn, J. Enzymatic Surface Erosion of High Tensile Strength
Maeda, M. Degradation and Adsorption Characteristics of PHB Polycarbonates Based on Natural Phenols. Biomacromolecules 2014,
Depolymerase As Revealed by Kinetics of Mutant Enzymes with 15, 830−836.
Amino Acid Substitution in Substrate-Binding Domain. Biomacromo- (514) Pastorino, L.; Pioli, F.; Zilli, M.; Converti, A.; Nicolini, C.
lecules 2010, 11, 113−119. Lipase-catalyzed degradation of poly(ε-caprolactone). Enzyme Microb.
(496) Shinomiya, M.; Iwata, T.; Doi, Y. The adsorption of substrate- Technol. 2004, 35, 321−326.
binding domain of PHB depolymerases to the surface of poly(3- (515) Kobayashi, S.; Uyama, H.; Takamoto, T. Lipase-Catalyzed
hydroxybutyric acid). Int. J. Biol. Macromol. 1998, 22, 129−135. Degradation of Polyesters in Organic Solvents. A New Methodology
(497) Mukai, K.; Yamada, K.; Doi, Y. Kinetics and Mechanism of of Polymer Recycling Using Enzyme as Catalyst. Biomacromolecules
Heterogeneous Hydrolysis of Poly[(R)-3-Hydroxybutyrate] Film by 2000, 1, 3−5.
Pha Depolymerases. Int. J. Biol. Macromol. 1993, 15, 361−366. (516) Sivalingam, G.; Chattopadhyay, S.; Madras, G. Solvent effects
(498) Doi, Y.; Kanesawa, Y.; Kunioka, M.; Saito, T. Biodegradation of on the lipase catalyzed biodegradation of poly (ε-caprolactone) in
microbial copolyesters: poly(3-hydroxybutyrate-co-3-hydroxyvalerate) solution. Polym. Degrad. Stab. 2003, 79, 413−418.
and poly(3-hydroxybutyrate-co-4-hydroxybutyrate). Macromolecules (517) Banerjee, A.; Chatterjee, K.; Madras, G. Effect of solvents on
1990, 23, 26−31. the enzyme mediated degradation of copolymers. Mater. Res. Express
(499) Chanprateep, S.; Shimizu, H.; Shioya, S. Characterization and 2015, 2, 095301.
enzymatic degradation of microbial copolyester P(3HB-co-3HV)s (518) Matsumura, S. Enzyme-catalyzed synthesis and chemical
produced by metabolic reaction model-based system. Polym. Degrad. recycling of polyesters. Macromol. Biosci. 2002, 2, 105−126.
Stab. 2006, 91, 2941−2950. (519) Osanai, Y.; Toshima, K.; Matsumura, S. Enzymatic trans-
(500) Panayotidou, E.; Baklavaridis, A.; Zuburtikudis, I.; Achilias, D. formation of aliphatic polyesters into cyclic oligomers using enzyme
S. Nanocomposites of Poly(3-hydroxybutyrate)/Organomodified packed column under continuous flow of supercritical carbon dioxide
Montmorillonite: Effect of the Nanofiller on the Polymer’S with toluene. Sci. Technol. Adv. Mater. 2006, 7, 202−208.
Biodegradation. J. Appl. Polym. Sci. 2014, 132, 1. (520) Kondo, R.; Toshima, K.; Matsumura, S. Lipase-catalyzed
(501) Murase, T.; Suzuki, Y.; Doi, Y.; Iwata, T. Nonhydrolytic selective transformation of polycaprolactone into cyclic dicaprolactone
Fragmentation of a Poly[(R)-3-hydroxybutyrate] Single Crystal and its repolymerization in supercritical carbon dioxide. Macromol.
Revealed by Use of a Mutant of Polyhydroxybutyrate Depolymerase. Biosci. 2002, 2, 267−271.
Biomacromolecules 2002, 3, 312−317. (521) Jbilou, F.; Dole, P.; Degraeve, P.; Ladavière, C.; Joly, C. A
(502) Numata, K.; Abe, H.; Doi, Y. Enzymatic processes for green method for polybutylene succinate recycling: Depolymerization
biodegradation of poly(hydroxyalkanoate)s crystals. Can. J. Chem. catalyzed by lipase B from Candida antarctica during reactive
2008, 86, 471−483. extrusion. Eur. Polym. J. 2015, 68, 207−215.
(503) Timmins, M. R.; Lenz, R. W.; Hocking, P. J.; Marchessault, R. (522) Zhu, X.; Fryd, M.; Wayland, B. B. Kinetic-mechanistic studies
H.; Fuller, R. C. Effect of tacticity on enzymatic degradability of of lipase-polymer micelle binding and catalytic degradation: Enzyme
poly(beta-hydroxybutyrate). Macromol. Chem. Phys. 1996, 197, 1193− interfacial activation. Polym. Degrad. Stab. 2013, 98, 1173−1181.
1215. (523) Ż enkiewicz, M.; Richert, A.; Malinowski, R.; Moraczewski, K. A
(504) Lenglet, S.; Li, S.; Vert, M. Lipase-catalysed degradation of comparative analysis of mass losses of some aliphatic polyesters upon
copolymers prepared from ε-caprolactone and dl-lactide. Polym. enzymatic degradation. Polym. Test. 2013, 32, 209−214.
Degrad. Stab. 2009, 94, 688−692. (524) Liu, F.; Zhao, Z.; Yang, J.; Wei, J.; Li, S. Enzyme-catalyzed
(505) Zhao, Z.; Yang, L.; Hu, Y.; He, Y.; Wei, J.; Li, S. Enzymatic degradation of poly(l-lactide)/poly(ε-caprolactone) diblock, triblock
degradation of block copolymers obtained by sequential ring opening and four-armed copolymers. Polym. Degrad. Stab. 2009, 94, 227−233.
polymerization of l-lactide and ε-caprolactone. Polym. Degrad. Stab. (525) Dong, J.; Liao, L.; Ma, Y.; Shi, L.; Wang, G.; Fan, Z.; Li, S.; Lu,
2007, 92, 1769−1777. Z. Enzyme-catalyzed degradation behavior of l-lactide/trimethylene

884 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885
Chemical Reviews Review

carbonate/glycolide terpolymers and their composites with poly(l-


lactide-co-glycolide) fibers. Polym. Degrad. Stab. 2014, 103, 26−34.
(526) Tsuji, H.; Ikarashi, K. In vitro hydrolysis of poly(l-lactide)
crystalline residues as extended-chain crystallites. Polym. Degrad. Stab.
2004, 85, 647−656.
(527) Zhao, Z. X.; Yang, L.; Hua, J. J.; Wei, J.; Gachet, S.; El Ghzaoui,
A.; Li, S. M. Relationship between enzyme adsorption and enzyme-
catalyzed degradation of polylactides. Macromol. Biosci. 2008, 8, 25−
31.
(528) Stürzel, M.; Mihan, S.; Mülhaupt, R. From Multisite
Polymerization Catalysis to Sustainable Materials and All-Polyolefin
Composites. Chem. Rev. 2016, 116, 1398−1433.

NOTE ADDED AFTER ASAP PUBLICATION


This paper was published to the Web on October 19, 2017,
with errors in Figure 3 and Scheme 3. These were corrected in
the version published to the Web on October 24, 2017.

885 DOI: 10.1021/acs.chemrev.7b00329


Chem. Rev. 2018, 118, 839−885

Potrebbero piacerti anche