You are on page 1of 22

Review

Applying the Principles of Green Chemistry to


Polymer Production Technology
,* Somaieh Salehpour
Marc A. Dube

The 12 principles of green chemistry are reviewed and applied specifically to polymer
production. Examples of how the principles relate to current practice in polymer reaction
engineering and which areas show the greatest potential impact for implementation of these
principles are discussed. This paper does not attempt to
be exhaustive but rather to target specific areas for
further development.

1. Introduction

(2)

Sustainability refers to the ability to meet current needs


without putting in jeopardy the ability of future generations to meet their own needs. The development of
sustainable technologies has been dealt with in several
ways, one of which is the application of the principles of
green chemistry to various processes. As a mature and
versatile field, the polymer industry plays a significant role
in our society as polymers have become ubiquitous. Issues
with the extensive use of fossil-based raw materials and
large amounts of reagents that are of environmental
concern, in addition to the accumulation of polymeric
materials in the environment, gives scientists and engineers the responsibility to re-examine the polymerization
process in light of the 12 principles of green chemistry.[1]
The 12 principles are:
(1)

Prevent waste: design chemical syntheses to prevent


waste, leaving no waste to treat or clean up.
, Dr. S. Salehpour
Dr. M. A. Dube
Department of Chemical and Biological Engineering, Centre for
Catalysis Research and Innovation, University of Ottawa, Ottawa,
ON, Canada
E-mail: marc.dube@uottawa.ca

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

(3)

(4)

(5)

(6)

(7)

Design safer chemicals and products: design chemical


products to be fully effective, yet have little or no
toxicity.
Design less hazardous chemical syntheses: design
syntheses to use and generate substances with little or
no toxicity to humans and the environment.
Use renewable feedstock: use raw materials and
feedstock that are renewable, rather than depleting.
Renewable feedstocks are often made from agricultural products or are the wastes of other processes;
depleting feedstocks are made from fossil fuels
(petroleum, natural gas, or coal) or are mined.
Use catalysts, not stoichiometric reagents: minimize
waste by using catalytic reactions. Catalysts are
used in small amounts and carry out a single reaction
many times. They are preferable to stoichiometric
reagents, which are used in excess and work only
once.
Avoid chemical derivatives: avoid using blocking or
protecting groups or any temporary modifications if
possible. Derivatives use additional reagents and
generate waste.
Maximize atom economy: design syntheses so that
the final product contains the maximum proportion of
the starting materials. There should be few, if any,
wasted atoms.

wileyonlinelibrary.com

DOI: 10.1002/mren.201300103

, S. Salehpour
M. A. Dube
www.mre-journal.de

(8)

Use safer solvents and reaction conditions: avoid


using solvents, separation agents, or other auxiliary
chemicals. If these chemicals are necessary, use
innocuous chemicals. If a solvent is necessary, water
is a good medium as well as certain eco-friendly
solvents that do not contribute to smog formation or
destroy the ozone.
(9) Increase energy efficiency: run chemical reactions at
ambient temperature and pressure whenever possible.
(10) Design chemicals and products to degrade after use;
design chemical products to break down into innocuous substances after use so that they do not
accumulate in the environment.
(11) Analyze in real time to prevent pollution: include inprocess real-time monitoring and control during
syntheses to minimize or eliminate the formation
of by-products.
(12) Minimize the potential for accidents; design chemicals and their forms (solid, liquid, or gas) to minimize
the potential for chemical accidents including explosions, fire, and release to the environment.
Given a general acceptance of these principles, we
herewith examine each principle as applied to current
practices in polymer reaction engineering.

1.1. Prevent waste: design chemical synthesis to


prevent waste, leaving no waste to treat or clean up
The generation of waste in polymer production arises
primarily from three sources: residual monomers due to
incomplete reaction conversion, waste from chemicals used
during the process but not consumed in the reaction such as
solvents and suspending agents, and off-spec material
generated due to excursions from controlled conditions.
Residual monomers are especially problematic because
of their typically hazardous nature. Since most monomers
demonstrate significant toxicity to human health, reducing
the residual monomer content is desired to prevent
workplace exposure as well as exposure to the consumer.
Residual monomer removal techniques are generally
classified into two categories: chemical methods and
physical methods. Chemical methods include the reaction
of the residual monomer to generate additional polymer or
to produce non-toxic or at least, easily removable compounds. Physical methods include the removal of residual
monomer from the polymer by evaporation, by solvent
extraction or with the aid of an ion-exchange resin.[2]
Extraction of the residual monomer using supercritical
carbon dioxide was reported to give better performance
than steam stripping techniques.[3]
Commonly used chemical methods include ramping the
reaction temperature and/or using a finishing catalyst

 has over 25 years experience in


Marc A. Dube
polymer reaction engineering as well as several
years in biodiesel production technology. His
expertise is focused on the study of reaction
kinetics, polymer characterization and the use of
on-line sensors for coatings, resins and adhesive
applications. Most recently, his work has focused
on sustainable practices in polymer production. He
has served as Chair of the Department of Chemical
and Biological Engineering and is a founding
member of the Centre for Catalysis Research and
Innovation (CCRI) at the University of Ottawa in
Canada. He is both a Fellow of the Chemical
Institute of Canada (FCIC) and a Fellow of the
Engineering Institute of Canada (FEIC).
Somaieh Salehpour obtained her Ph.D. degree and
M.A.Sc degree in Chemical Engineering at the
University of Ottawa in Canada under the
. Her research work
supervision of Dr. Marc A. Dube
focused on polymer syntheses using renewable
and environmentally friendly monomers and
solvents. Currently, she is a Senior Research
Engineer at EcoSynthetixInc, where her work
focuses on development and manufacturing of
wide range of new environmentally friendly
products with enhanced performance, cost benefits and an improved environmental footprint
when compared to the petroleum-based products
they replace.

often referred to as a chaser. This protocol is usually


implemented immediately after the polymerization and
may lead to additional costs due to the extra processing
stage. Post-polymerization methods and/or chemical
monomer removal are often employed to reduce the
monomer content before the use of devolatilization
processes. Other techniques of this type include postcatalysis procedures followed by spray-drying,[4] and
hydrolytic slitting of the monomers followed either by
distillation or by the use of an oxidizing agent.[5]
In addition to the above-mentioned methods, there are
ways to ensure that the monomer consumption in
the polymerization is as high as possible. An example is
the formation of poly(styrene) (PS) latex via ultrasonic
initiation.[6] The presence of ultrasonic irradiation results in
the increase of the polymerization rate. The molecular
weight of the polymer latex obtained from this method is
higher than the conventionally manufactured one, yielding
a higher monomer conversion and lower residual monomer
content in the final polymer product.[6] Modifying the
reaction formulation by changing some of the components
(e.g., initiator type) can also give promising results.[7]
Similarly, optimization of process conditions can also lead
to significant improvements to residual monomer levels, as
shown for the emulsion polymerization of vinyl acetate
(VAc).[8]

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Applying the Principles of Green Chemistry to Polymer . . .


www.mre-journal.de

Choosing the best residual monomer removal technique is highly case specific as each polymerization
method and monomer system presents its own peculiarities. However, one major principle that governs the
choice of residual monomer techniques is that the
method used should not affect the final properties of
the polymer product.
Solution polymerization is a widely used polymerization technique, which uses a solvent as the polymerization medium. Solution polymerization is often preferred
over bulk polymerization due to viscosity and heat
transfer issues.[6] One major disadvantage of solution
polymerization is the removal of the solvent from the
polymer, which may require energy intensive methods
such as distillation.[6] However, not all polymerizations
can be practically carried out in a solvent-free environment, and many solution polymerization products
exhibit superior properties. Therefore, many research
efforts are underway to achieve similar properties
in either alternative, safer solvents or by other polymerization techniques.[911]
The issue of off-spec material concerns many industries.
In the polymer industry, the production of off-spec
material is not unusual. The viscous nature of high
molecular weight polymers and the typically exothermic
nature of the reaction make the polymerization process
prone to temperature excursions. These excursions
can affect the polymer molecular weight and lead to
branching, cross-linking and eventually, the formation of
gel. In some cases, the presence of cross-linked materials
and gels is desired to impart certain performance properties. However, in many instances, these may have a
deleterious effect on product properties. Ensuring proper
mixing, heat removal and process control methods can be
helpful in those cases.
In addition, off-spec material can result from the
presence of impurities found, e.g., in the monomer feed
stream. Impurities that scavenge radicals in free-radical
polymerizations can have a significant effect on the final
product.[12] One should also consider that bad batches are
often blended with good ones. Finally, recent advances in
polymerization methods such as controlled radical polymerization (CRP) have served to restrain the formation of
unwanted compounds and have also enabled remarkable
microstructural control.[1315]
As can be deduced from the above discussion, there is not
a single pot solution for the minimization of waste
production in polymerization processes. Often, a combination of several steps must be considered. These steps include
measures taken to reduce the use of solvents or their
replacement, the implementation of catalysis to improve
yield, and means to either prevent or reduce residual
monomer. The use of alternatives to current solvent-based
synthesis may play a significant role in reducing waste, but

www.MaterialsViews.com

the maintenance of high quality final product properties is


of concern.
1.2. Design safer chemicals and products: design
chemical products to be fully effective, yet have
little or no toxicity
In a relative way, polymer products tend to be inherently
safer than their starting materials (e.g., monomers,
solvents). In general, most polymers are in and of
themselves non-toxic. However, in the event that they
are degraded or burned, their degraded components may
present serious environmental effects.[1618] In addition,
incomplete conversion of the starting materials may result
in residual monomers in the product, which upon release,
can prove to be quite harmful.[1618] Moreover, in many
instances, a variety of additives are often used to modify
polymer properties and these also can pose health and
environmental risks.[19,20] These additives can be used as
processing aids, antioxidants, flame retardants, solvents,
electrostatic agents, stabilizers, pigments, mechanical
property modifiers (e.g., plasticizers), etc.[20]
Solvents are a major input in the polymerization industry
and comprise a significant fraction of the waste generated
as well as a large portion of workplace hazards. A significant
source of toxicity in polymer products occurs when solvents
are used either as diluents or to enable the application of
the polymer (e.g., as a paint or adhesive). Volatile organic
compounds (VOCs), highly polar conventional solvents
such as N-methylpyrrolidone, N,N0 -dimethylacetamide, N,
N0 -dimethylformamide, pyridine, and chlorinated solvents,
which are used in condensation polymerizations, result in
increased air pollution, most of them being volatile and
most of them toxic and flammable.[2123] These solvents are
high on the list of harmful chemicals because they are used
in significant volumes and their volatility makes them
difficult to handle. Hence, based on these concerns, certain
less hazardous surrogates have been suggested for use.
Among the suggested solvent alternatives are monoterpenes (MTs), ionic liquids (ILs), and supercritical fluids
(mainly, supercritical carbon dioxide), to name but a
few.[2123] These environmentally benign substituent
solvents perform comparable to conventional fossil-fuel
based solvents.[2123] A more in depth discussion on solvent
alternatives is presented in Section 8.
Functional fillers represent a significant component of
polymer products.[24] They are primarily used to modify
mechanical and thermal properties. Many synthetic fillers
such as silica, titanium dioxide, calcium carbonate, glass
fiber, and carbon fiber are used extensively. An increasing
number of naturally sourced fillers such as clays, natural
fibers (e.g., cellulose, hemp, flax, kenaf, and wood) and
starch are also employed. Attempts to address environmental concerns regarding these composite materials have

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

, S. Salehpour
M. A. Dube
www.mre-journal.de

been reviewed recently.[25,26] Some specific examples


include the use of pine wood flour,[27] natural jute fiber,[28]
sugarcane bagasse fiber,[29] cellulosic fibers,[30] and kenaf
biofibers.[31]
An interesting subclass of fillers, nanofillers, are materials having at least one dimension in the nanometer range.
Due to their high aspect ratio and high surface area, even
low loadings of these nanomaterials impart unusually
strong modifications to polymer properties.[32] Whether
synthetic or natural, when these materials are modified to
the nanoscale, it is not known how these might impact the
environment and human health. At least one of these,
cellulose nanocrystals, has been approved in some sectors
as environmentally benign.[33]
Plasticizers are incorporated into a material to increase
its flexibility and processability.[34] These are generally nonvolatile, low molecular weight materials, which consequently, can leach out of the polymer product. Phthalates,
which are the diesters of 1,2-benzenedicarboxylic acid
(phthalic acid), are the most common class of plasticizers
used in industry. Others include benzoates, tartrates,
chlorinated hydrocarbons, phosphates, etc.
Other examples of toxic additives used in plastic
processing for which their impacts on human health have
been extensively discussed[19] are bisphenol A and polybrominated diphenyl ethers (PBDEs). Bisphenol A is used in
a variety of polymer products mostly as food can lining[35]
and also in poly(carbonate) plastics as used for baby bottles
and other containers.[36] PBDEs are flame retardant
chemicals used in many thermoplastics. Studies have
shown that exposure to bisphenol A and PBDEs causes
altered endocrine function and reproductive effects[37] and
might also result in cancer in the case of higher percentage
accumulation.[38]
Bisphenol A has been utilized in four main classes of
polymers: epoxies, polycarbonates, polyesters, and polyimide. In order to minimize the level of this endocrinedisrupting additive, several pathways and replacements
have been widely investigated.[39] For example, aliphaticaromatic components derived from 2,2,4,4-tetramethyl1,3-cyclobutanediol (TMCBD)[40] and isosorbide have been
studied as promising candidates for bisphenol A
replacement.[41]
Another class of highly toxic chemicals which have been
used for cross-linking and network formation are highly
reactive isocyanate containing compounds.[42] Isocyanates
are used extensively in polyurethane production resulting
in outstanding mechanical and thermal properties.[43]
Replacement of isocyanate components with other, less
harmful di- and tri-functional molecules such as bispropargyl-succinate has been reported.[44] Additionally,
less toxic routes to the synthesis of isocyanate-free
polymers or non-isocyanate polyurethanes (NIPUs) have
been explored using different monomers. For example,

10

monomers such as cyclic carbonates, for the synthesis of


poly(urethane-carbonate)[45] and carbonated soybean oil
for the catalytic synthesis of NIPUs have been studied.[46]
A logical approach has been to replace these compounds
with higher molecular weight materials to achieve lower
solubility and migration rates. Research on a number of
natural-based plasticizers is underway.[47] Notable examples include soybean oil, epoxidized triglycerides from
vegetable oils, water, polyols (e.g., glycerol, sorbitol, and
xylitol), monosaccharides (e.g., mannose, glucose, fructose,
and sucrose), amino acids, saturated fatty acids. The use of
triethyl citrate in the development of a cellulose acetate
biopolymer is a good example of combining a renewable,
non-toxic plasticizer with a renewable polymer.[48]
The presence of VOCs as solvents, residual monomers,
and additives in the polymer industry poses many potential
negative consequences to human health and the environment.[19,49,50] The replacement or removal of these components without substantially affecting the quality and
properties of the polymer products is an area of ongoing
study. The use of alternative solvents as suggested in
Section 8 presents an area of significant impact. The
presence of residual monomers, while not as significant in
quantity as other volatile components, nonetheless,
provides an important area for improvement, especially
in light of the increased use of polymers in medical
applications. Polymer additives have only begun to be
targeted by various concerned groups and their replacement with environmentally benign compounds is also at
the forefront of several research efforts.
1.3. Design less hazardous chemical synthesis:
design syntheses to use and generate substances with
little or no toxicity to humans and the environment
The transformation of one or more monomers into a final
polymer product often includes a number of additional
components to aid in the synthesis or to modify the final
product properties. As noted in the previous section, large
volume of solvents, often petroleum-based, as well as
plasticizers, fire retardants, pigments, stabilizers, catalysts,
and the like, are used throughout the synthesis stage and
these often pose serious toxicity concerns. In addition to the
workplace hazards posed by these materials, their release to
the environment during product use, under unexpected
conditions (e.g., in a fire), and after disposal may also pose
significant risks (see Section 2). Hence, a less toxic polymer
synthesis suggests either an alternative processing technology or the replacement of toxic starting materials with
environmentally benign substances (see also Sections 4, 5,
6, and 8). The discussion that follows is focused on the use of
less toxic monomers, solvents, catalysts, and additives as
possible surrogates for conventional components in
polymer product synthesis, processing, and modification.

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Applying the Principles of Green Chemistry to Polymer . . .


www.mre-journal.de

It is widely accepted that many monomers pose serious


health hazards but in a polymerized form, the hazard is no
longer present or is significantly reduced. Of course, the
hazard of exposure to monomers does exist during polymer
synthesis in the workplace but may also exist in the product
in the event of incomplete conversion, where residual
monomer remains. One also should not forget cases where
the curing of the polymer must take place (e.g., in epoxies).
Eventually, depending on the usage, degradability, and
disposal methods, a polymer may revert to its original
monomer components and once again, the toxicity of the
monomer would become important. Although there are
synthetic polymers that are not bio-sourced and still
environmentally benign, a shift to the use of so-called
green monomers often implies the use of starting
materials from biomass.
Bio-based polymers, like polymers from cellulose and
ligno-cellulosic materials, were reported to present poorer
performance compared to those derived from petroleumbased monomers, especially in terms of thermal processability.[51] The relative difficulty of converting biomass
into plastics is due to celluloses rigid backbone. This is a
common refrain when it comes to replacing long-existing
products with more benign options. It is clear that work has
to be done to improve the performance of bio-based
materials but successes are indeed possible.[52,53] Below,
we report on the numerous examples of bio-sourced
monomers being successfully transformed to quality
polymer products.
Cellulose derivatives are attractive bio-based materials,
among which nitrocellulose (NC) and cellulose acetates (CA)
are the most common. Many recent studies have focused on
CA, mostly because it is produced in large quantities.[54]
Lignocellulosic material, primarily wood, is another class
of biomass that can be processed into plastics. However,
lignocellulosic materials cannot flow easily and methods
for processing them are limited. Several reactions such as
esterification and etherification were reported during the
late 1970s and early 1980s to convert wood into a plastic
material by chemical modification.[55] Such a conversion
was mainly related to the internal plasticization of wood,
and the degree of plasticization was found to be insufficient,
especially when small substituent groups or polar groups
were used. Therefore, external plasticization was used to
achieve better properties. Benzylated wood (BzW) is one of
the most common externally plasticized lignocellulosic
materials with reported tensile strengths of 42.7 MPa,
which are significantly higher than that of PS at 29.4 MPa
under the same conditions.[51] In addition, differences are
observed in their flow properties, with PS starting to flow at
153 8C whereas BzW begins to flow at 175 8C.[56]
Poly(lactic acid) (PLA) is another class of bio-based
polymers which is highly biocompatible; not surprising
given that its monomer, lactic acid, is a natural product

www.MaterialsViews.com

found in the body.[5760] Early application of PLA was in the


biomedical field[5860] and greater improvements in its
synthesis have been achieved over the past decade,
particularly from fermentation pathways to convert corn
starch into lactic acid. PLA is used in a wide range of
applications such as packaging materials, plastic bags, and
thermoforms.[61,62]
Biodegradable poly(carbonate) synthesized by ring
opening polymerization of six-membered cyclic carbonates
is another interesting class of polymers that can be used for
the production of thermoplastics. These polymers are
widely used in the medical field, e.g., in the production of
sutures, drug delivery systems, implants, and tissue
engineering.[6366]
Several other benign monomers exist either as naturally
occurring compounds or through certain chemical modifications based on natural substances. Conversion of starch
into environmentally benign monomers via catalytic
reduction has been studied and compounds such as polyols,
organic acids (lactic and glycolic acid), 5-hydroxymethylfurfural, and levulinic acid have been formed as a result of
the catalytic breakdown of starch.[67]
Glycerol, which can be obtained via various sources, such
as biodiesel production, is another interesting renewable
source for the production of poly(glycerol).[52] Stimuliresponsive hydrogels were produced from bio-based poly
(glycerol).[53] A wide variety of polyurethane materials have
been produced from different types of vegetable oil-based
monomers. These polyurethanes have been evaluated and
proven to be of comparable, or even better performance
than some fossil-based polymers.[68]
In order to modify the final properties of polymer
products, several additives may be added such as emulsifier,
plasticizers (see Section 2), stabilizers, flame retardants, and
other fillers (see Section 2). For example, nonylphenol
ethoxylates (NPEOs) are extensively used as surfactants for
emulsion polymerization and for stabilization in latex
applications. But NPEOs are undesirable because alkylphenol residues are formed when these surfactants are
released to the environment.[69] Therefore, the substitution
of ionic (NPEOs) with environmentally friendly non-ionic
alkyl phenol free emulsifiers is highly encouraged.[70] Fatty
alcohol ethoxylates (FAEOs) as less toxic surfactants are
suitable substitutes for these surfactants. FAEOs, having a
degree of ethoxylation between 7 and 40, show excellent
water miscibility and stability performance in emulsion
polymerization.[70] The hydrophobe in the surfactants is
synthesized from renewable resources and the use of these
bio-based surfactants has proven to be successful for many
systems such as acrylates.[70] Other examples of environmentally friendly surfactants are mannuronate moieties
derived from alginates, fatty hydrocarbon chains derived
from vegetable resources[71] and ILs[72] and surfactants
based on lipo-peptides.[73]

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

11

, S. Salehpour
M. A. Dube
www.mre-journal.de

Flame retardants are also common additives used


especially in the plastics industry. Two significant classes
of flame retardants are organohalogen and organophosphorous compounds.[7476] The organohalogen compounds,
particularly brominated aromatics, have been the most
widely used because of their efficiency and low cost.
Alternatively, organophosphorous compounds have equal
efficiency to the organohalogen compounds, but are much
more expensive. The bioaccumulation of organohalogen
compounds in the environment and the related health risks
are significant concerns.[77,78] Recent research has been
conducted on a non-toxic flame retardant derived from
tartaric acid, a by-product of the wine industry, and can be
used in the synthesis of bio-based, environmentally
friendly flame retardants for polymers.[79] One can also
look to bio-based nanofillers for improved flame retardant
properties.[32]
Commercial polymers are often processed in the
presence of solvents. Application of environmentally
friendly solvents in polymer production is discussed
extensively in Section 8. Less hazardous syntheses and
the use of substances with little or no toxicity to humans
and the environment is one of the most important areas of
research in polymer science and many current studies are
focused on this field. This is not surprising as there is great
potential for significant impact on harm reduction from
these approaches.
1.4. Use renewable feedstock: use raw materials and
feedstock that are renewable rather than depleting.
Renewable feedstocks are often made from
agricultural products or are the wastes of other
processes; depleting feedstocks are made from fossil
fuels (petroleum, natural gas, or coal) or are mined
The use of macromolecular materials based on renewable
resources had long been a low priority with the surge of
coal-based chemistry starting in late 19th century, followed
by the petrochemical revolution of the 20th century.[80]
However, in recent years, increasing environmental concerns and the need for more versatile polymer-based
materials has led to greater interest in polymers from
renewable feedstocks. This includes the design and
synthesis of renewable polymers with various macromolecular architectures and enhanced properties.[81] The
consumption of fossil fuels in the production of plastics
was reported to be about 7% of worldwide oil production.[82]
Therefore, given the increasing oil prices and its limited
sources, it is necessary to develop new pathways to
polymeric materials using renewable resources. Some
interesting renewable sources are discussed below.
The use of renewable monomers can be attractive due to
their abundance, the typical presence of high energy
heteroatom functionality, and in many cases, low cost.

12

Challenges may arise due to the frequent need to eliminate


superfluous heteroatom functionality in these monomers
and in some cases, high cost. It is also frequently difficult to
achieve high monomer purity from bio-based sources and
therefore, high molecular weight polymer production can
be hampered.
Gandini[83] classified renewable monomers from the
following starting materials (limited examples are shown):
(1)

vegetable oils: soybean oil, castor oil, palm oil,


rapeseed oil, linseed oil, tung oil, etc.
(2) lignin fragments: cellulose, chitin, chitosan, starch,
lactic acid, etc.
(3) sugars
(4) rosin: resin acids
(5) glycerol
(6) furans: furfural, furfuryl alcohol
(7) tannins
(8) suberin
(9) terpenes: a-pinene, b-pinene, and limonene
(10) and others: citric acid, tartaric acid, carbon dioxide
Lactic acid is a product of the fermentation of D-glucose,
usually derived from corn or sugar feedstocks. The lactic
acid is thermally and catalytically converted into its cyclic
dimer and lactide. The lactide can then undergo ringopening polymerization to yield PLA (a.k.a. poly(lactide) or
PLA).[61] PLA makes an interesting replacement for polyolefins, yet degrades to metabolites. It has found extensive
application in the packaging and fiber industries.[61]
Plant oils, such as soybean oil, palm oil, and rapeseed oil
are some of the most important renewable raw materials
for the chemical industry. Plant oils consist of a broad
composition of fatty acids (i.e., triglycerides) and are
typically liquid at room temperature. Their physical and
chemical properties are greatly affected by the stereochemistry of the double bonds of the fatty acid chains, their
degree of unsaturation, and the length of the carbon chain
of the fatty acids. Plant oils find broad application in foods,
pharmaceuticals, cosmetics, fuels (e.g., biofuels), lubricants,
paints, and construction materials.[84] Over 120 million tons
are produced per year worldwide, with 12 million tons
available to the oleochemical industry; the rest is used for
food. Due to their natural abundance and reactive
functionality, they make an obvious choice as polymer
building blocks. They can be classified into two product
groups: base chemicals (e.g., fatty acids, methyl esters, fatty
alcohols) and derivative compounds. Plant oil-based
polymers have found extensive use in paints, coatings,
resins, and as flooring materials. The most famous example
of these is linoleum flooring, which has been industrially
produced for well over 100 years.[85] Linoleum provides a
durable and environmentally friendly alternative to PVC
flooring. In recent years, extensive studies have been carried

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Applying the Principles of Green Chemistry to Polymer . . .


www.mre-journal.de

out on the production of various polymers from plant


oils.[8690] However, their use as monomers presents two
major challenges: (i) their chemical structure is often
heterogeneous and variable; and (ii) extraction of plant oils
can be costly due to their typically low concentrations and
seasonally variable compositions.[82]
Carbon dioxide (CO2) is an attractive monomer replacement option since it is abundant, inexpensive, nonflammable, and a waste product of many chemical
processes. It is estimated that over 200 billion tons of
glucose are produced from CO2 via photosynthesis each
year. However, the development of effective processes that
use CO2 is lagging.[82] For example, the viable production of
only urea, salicylic acid, and some cyclic carbonates are
possible at this time. Nonetheless, interesting copolymerization products are possible by combining CO2 with
heterocycles (e.g., epoxides, aziridines, and episulfides).
On the larger scale, CO2 can be used as a feedstock for the
synthesis of poly(carbonate).[91] Poly(carbonate) is a
thermoplastic material and can be efficiently produced
from CO2 and epoxides.
Due to a significant glut in the glycerol market because of
increasing biodiesel production, there is great interest in
finding economic ways to utilize the emerging surplus of
glycerol. Glycerol as a bio-based monomer with high
functionality is an interesting candidate for sustainable
step-growth polymerization. Recently, new opportunities
for the conversion of glycerol to high molecular weight
polymers have emerged.[52] Poly(glycerol) produced from
this cost-effective renewable monomer was used to
produce hydrogels suitable for a wide range of pharmaceutical and biomedical applications.[53,92]
Renewable bio-based aromatic monomers have been
synthesized and used as alternatives to monomers such
as terephthalic acid and styrene which are widely used
fossil-based aromatics.[93] Efficient conversion of solid
biomass feedstock to high quality aromatics has been
investigated using different catalytic methods such as
catalytic fast pyrolysis.[94] From these, one can derive furan
heterocycles and other similar molecules.[95] Conversion of
the furan aglycon into g-keto-carboxylic acid moieties
results in a renewable building blocks for different
applications.[96] In another approach, furan moieties
were reacted with epoxidized triglycerides, another monomer from two renewable sources, in a DielsAlder reversible
polycondensation.[97] Additional discussions on bio-based
aromatics (e.g., terpenes) can be found in other sections.
In addition to renewable monomers, natural polymers
such as starch can be used as renewable feedstocks.[98]
Starch can be found in nature in different plant forms and
this high molecular weight polymer can be modified to be
used as a replacement for fossil-based polymers. For
example, starch was used as the starting material for
production of EcoSphere biolatex binder which has been

www.MaterialsViews.com

proven to be a successful replacement for petrochemicalbased latexes such as styrene butadiene rubber (SBR)
used in the manufacture of paper and paperboard
products.[99,100]
Research in the use of renewable monomers is expanding
rapidly and one cannot hope to provide an exhaustive
review of each of the possibilities. From terpenes[101,102] to
furans[103] to carbohydrates,[104] and many others,[105,106]
interest in this area continues to grow. While current limits
to the feedstocks for renewable polymers imply that
they are not expected to overtake the market for commodity
polymers in the near future, the use of renewable
polymers in a wide range of specialty products is highly
promising.
1.5. Use catalysts, not stoichiometric reagents:
minimize waste by using catalytic reactions.
Catalysts are used in small amounts and carry out a
single reaction many times. They are preferable to
stoichiometric reagents, which are used in excess
and work only once
Catalytic agents are used to not only increase reaction rate
but to manipulate the final product properties. In free
radical polymerization, e.g., a small amount of catalyst say,
109103 M, results in the reaction of 0.110 M reactive
groups. Usually, by using catalysts, waste formation is
significantly decreased. In some cases, it is possible to
recover or recycle the catalyst after completion of reaction
but often separating the catalysts from the polymer product
is very challenging.[107] However, the catalysts are
often used at very low concentrations and can be
deactivated.[108,109]
The types of catalysts used in polymerization are either
homogeneous or heterogeneous. A majority of homogeneously catalyzed processes exist in the liquid phase.
Homogeneous catalysis is advantageous in terms of
increased activity but problems may arise due to corrosion
of reaction vessels and difficult separation processes.[109]
On the other hand, the use of heterogeneous catalysts has
proved to be an increasingly interesting alternative as
solid catalysts are rarely corrosive, are capable of withstanding a wide range of temperatures and pressures
without a significant impact on their activity, and
separation of the catalyst from the product may be
straightforward.[109]
Catalysts have been used in polymer synthesis since
1953, where some transition metal compounds (e.g., TiCL4)
were used in combination with aluminum alkyls for
polymerization of ethylene into linear poly(ethylene)
(PE).[110] The use of ZieglerNatta polymerization has
increased tremendously in the polymer industry over the
years for the production of polyolefins. By using Ziegler
Natta catalysts, it has been feasible to achieve outstanding

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

13

, S. Salehpour
M. A. Dube
www.mre-journal.de

balance between properties such as fluidity, impact


strength, and transparency. The production of polyolefins
via ZieglerNatta polymerization now exceeds 70 million
tons globally.[110] After some studies were performed on
titanocenes,[111] a group of catalysts containing homogeneous and single-site systems were developed, based on
metallocenes[112118] or late transition-metal complexes.[119,120] These catalysts seemed to be much better
than conventional ZieglerNatta systems. However, heterogeneous ZieglerNatta catalysts are still the most
popular catalyst in most industrial applications. Polymer
production technologies have improved to a great extent
with the evolution of ZieglerNatta catalysts. For example,
exceptional control of morphology over both the catalyst
and polymer particles was achieved and some important
polymer parameters such as molecular mass, molecular
weight distribution, and co-monomer incorporation were
obtained.[119,120]
Catalysts are also used in CRP methods such as atom
transfer radical polymerization (ATRP). This method, which
was originally established in the 1990s enables great
molecular weight and functionality control in polymerization with minimal catalyst concentrations. Copper complexes are the most popular catalysts used via this
method.[121123] Other metal catalysts are also used for
polymer syntheses[124] such as ruthenium catalyst for
living radical polymerization of star polymers[125], iron for
ATRP of methyl methacrylate (MMA),[126] and cobalt for
polycarbonates and cyclic carbonates syntheses.[127]
The concentration of catalysts used in polymerizations
can be an issue and, in many cases, residual catalyst in the
final polymer product could result in changes in final
properties as well as inducing toxicity. In previous decades,
advances in catalyst technology and properties such as
activity, selectivity, and specificity have led to the use of low
concentrations of catalysts with very high yield, which
minimizes the chance of occurrence of undesirable side
reactions.[110] For example, for PE synthesis, efficiencies of
catalysts based on either MgCl2 or its precursors, range from
approximately 10 to 40 kg polymer  g1 catalyst.[110]
During the past two decades, polymer synthesis using
environmentally benign enzymatic catalysis has received a
great deal of attention.[128] These catalysts are interesting
because they are reported to induce faster reaction rates,

result in efficient reactions at mild conditions and they have


high stereo-, regio-, and chemoselectivities.[129] However,
the use of enzymatic, bio-based catalysts, and organocatalysis is usually less straightforward than that of
chemical catalysis.[130,131] Their use often eliminates the
need for severe polymerization conditions such as elevated
temperatures for condensation polymerization, as enzymes
function under mild conditions (100 8C).[132] In addition,
precision syntheses are often achieved using enzymatic
catalysis. The synthesis of polyesters is a great example
of enzyme-catalyzed condensation polymerization at mild
conditions[132] and that of poly(ethylene glycol) using
Candida antarctica lipase B is an example of precision
polymerization.[133] Table 1 shows the different types of
polymers produced with their respective enzymes.[129]
In step growth polymerizations, the catalyst plays a
critical role. In addition, stoichiometric imbalances can
greatly and negatively affect the degree of polymerization.
The use of superacid catalysts for polydyroxyalkylation
reactions provides an example of overcoming this difficulty
and permitting the synthesis of polymers with a broad
range of molecular architecture.[134]
Catalysis is a highly evolved and advanced sector in
polymer reaction engineering and serves a key industrial
role in increasing polymerization rates and manipulating
final polymer product properties. For instance, most stepgrowth polymerizations require catalysts in order to
proceed at a practical rate of reaction. The introduction of
catalysts to such systems will effectively result in the
decrease in waste and minimization of the energy
consumption for running the reaction.[135]
1.6. Avoid chemical derivatives: avoid using
blocking or protecting groups or any temporary
modifications if possible. Derivatives use additional
reagents and generate waste
Throughout the polymerization process, there may be
instances in which the monomer must be modified either to
make it compatible for future steps (e.g., modification of
polymer surfaces) or to impart certain characteristics vital
to the polymer product. Such temporary or permanent
modifications are sometimes affected by the use of blocking
or protecting groups, which introduce additional reagents

Table 1. Examples of in vitro production of polymers catalyzed by enzymes.[132]

Enzymes

14

Typical polymers

Oxidoreductases

Polyphenols, polyanilines, vinyl polymers

Transferases

Polysaccharides, cyclic oligosaccharides, polyesters

Hydrolases

Polysaccharides, polyamino acids, polyamides, polyesters, polycarbonates


Macromol. React. Eng. 2014, 8, 728
2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Applying the Principles of Green Chemistry to Polymer . . .


www.mre-journal.de

into the process. These additional reagents may lead to the


generation of waste. For example, poly(vinyl alcohol) (PVA),
is usually modified temporarily using a t-butoxycarbonyl
group (t-BOC) into poly(vinyl-t-butyl carbonate) and then
thermolyzed to achieve the final PVA product. Such a
modification is done in order to make PVA soluble in a wide
variety of organic solvents, but during the process, carbon
dioxide and 2-methyl-propene are released, which are of
concern for workplace safety and the environment.[136]
Therefore, alternative means by which the desired product
performance properties can be achieved without the use of
additional reagents is important.
Another example is 2-hydroxyethyl methacrylate
(HEMA), a functional monomer which has been used in
many applications such as in the manufacture of soft
contact lenses.[137,138] HEMA homopolymer is a hydrophilic
gel synthesized through a fairly complicated pathway.
There have been several reports on the synthesis of
controlled-structure HEMA-based block copolymers via
anionic polymerization and in most of these approaches,
the protection of the alcohol functionality in the presence of
additional reagents is required.[139,140] These methods are
harmful to the environment due to the release of wastes. In
addition, the three-step synthesis method includes synthesis of the protected monomer followed by the controlled
polymerization and finally, the removal of the protecting
groups.
The application of blocking and protecting groups for
temporary modifications on polymeric substances is often
applied via CRP, but the outstanding property control
afforded by these methods presents an intriguing dilemma.
The ability to create unique polymer architectures under
highly controlled conditions would not be possible without
these methods. One form of CRP, ATRP, was discovered in
1995[121] and has been used effectively for the polymerization of hydrophilic monomers in aqueous solution under
mild conditions.[141143] For the production of poly(HEMA)
in aqueous media, the process was made more efficient by
switching the medium to a 50:50 methanol/water mixture
which resulted in a useful product without the use of a
protecting group to aid in solvation.[144] Similar results
were obtained when ATRP was employed to form
3-sulfopropyl methacrylate (SPMA) homopolymers and
amphiphilic block copolymers with MMA in a water/
dimethylformamide (DMF) mixed solvent at 20 8C using
Cu/bipyridine catalyst.[145] It is interesting that in polymerizations using a mixed solvent system, it is possible that
hydrophobic monomers can also be block-polymerized in
the same solvent. This allows for the direct synthesis of
hydrophobichydrophilic block copolymers without the
application of protecting group chemistry or post-polymerization derivatization. However, despite the lack of
protecting groups, there were some drawbacks. In particular, the small amount of residual SPMA monomer detected

www.MaterialsViews.com

during the initial stage of the MMA polymerization had a


significant effect on the micellization properties.[146]
Frequently, one finds the application of blocking and
protecting groups in the preparation of polymeric materials
for biomedical applications. In fact, blocking and protecting
groups are used widely in biomedical polymers because of
the need for bulk and surface modification. Such modifications are put in place to improve the acceptance and
healing of the synthetic device or medical implant. Many
innovative processing technologies are used to produce
functional devices resembling natural tissues.[146] Often,
the synthetic materials fabricated for medical purposes
need to be made compatible with reactions taking place in
the body, via the incorporation of reagents as blocking or
protecting groups.
The generation of waste that arises as a result of the
introduction of further reagents is undoubtedly an issue to
be dealt with in designing sustainable chemical processes.
In the case of polymer reaction engineering, the related
impacts are not as serious as some of the other issues
presented herein.[147,148] Many of the polymer products
requiring protecting/deprotecting are niche products. Thus,
given that worldwide production levels of these products
are significantly lower than that of commodity polymers
such as, e.g., PS or PE, the total impact of additional
chemicals used in their production is comparatively
small.
1.7. Maximize atom economy: design syntheses so
that the final product contains the maximum
proportion of the starting materials. There should be
few, if any, wasted atoms
Efficiency is a key factor in sustainable polymer
synthesis; it should be addressed not only in terms of
selectivity [chemo- (functional group differentiation),
regio- (orientational control of two reacting partners),
diastereo- (control of relative stereochemistry), and
enantio-selectivity (control of absolute stereochemistry)],
but also through the number of atoms in the reactants
(monomers and other reagents) that are retained in the
final polymeric product.[149] This concept is referred to as
atom economy, and is calculated by dividing the
molecular weight of the desired product by the sum of
the molecular weights of all the substances produced in
the stoichiometric equation. In ideal terms, a 100% atom
efficient process signifies one in which the product would
include all of the atoms of the reactants and no waste is
produced. The degree of atom economy is commonly
defined by the E-factor, the ratio of the amount of waste
produced to the amount of product that results (kg of
waste/kg of product).[150] Therefore, the lower the value of
the E-factor, the less waste is produced, and vice-versa.
Accordingly, the major benefits derived from an atom

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

15

, S. Salehpour
M. A. Dube
www.mre-journal.de

economical process are most notably the effective use of


valuable raw materials, decreased emissions, and low
waste production.
In the case of straightforward reactions, atom efficiency
can simply be enhanced by using catalysts. Using catalysts
facilitates the pathway by which a desired product can be
achieved by better selectivity and activity. Therefore, the
formation of unwanted by-products is minimized and if
possible, leads to a 100% atom economical process.[151]
However, in the case of polymerization, there is not any
one-pot solution to transform a less atom economical
process towards a more atom efficient one as there are
multiple factors responsible for the production of unwanted by-products. Atom wastage in polymerization, in some
instances, is unavoidable due to the nature of the process
and the absence of alternative ways to synthesize the
polymer. The most common example of polymerization in
which by-products are produced is found in many stepgrowth polymerizations. There is usually a specific type of
chemical reaction which enables monomers to combine to
form macromolecules. Example reactions in step-growth
polymerization include esterification, amidation, formation
of urethanes, and aromatic substitution.[135] Due to a limited
number of chemical reaction options, the polymer industry
has been forced to follow these pathways even though
unwanted by-products were continually being produced.
For instance, Nylon 66 is produced according to:[135]

As seen in the reaction scheme above, there is always one


mole of water produced as a by-product for every linkage
produced. Hence, this reaction is not fully atom efficient,
with an atom efficiency of 91.26%. On the other hand, there
are also 100% atom efficient step-growth polymerizations
such as in the production of polyurethanes, as shown in the
following reaction scheme:

Unlike most step-growth polymerizations, which proceed via a single pathway, there are cases such as in the
formation of poly(ethylene terephthalate) (PET), which can
proceed via several pathways. PET can be formed in two
distinct ways:

16

The first pathway is moderately more atom efficient than


the second, where atoms are lost as 2n-H2O. Based on this,
one may decide to proceed with the first pathway in view of
the atom economy, but availability of the corresponding
monomers, price, reaction conditions, energy consumption,
and safety also must be considered.[135] A similar scenario
can be witnessed in the production of poly(glycerol), which
can be synthesized via two pathways, i.e., (i) using glycidol
as a monomer via free radical polymerization, which has a
higher atom economy and (ii) from glycerol via step-growth
polymerization, which has a relatively lower atom economy.[52,53] In the first case, a highly toxic, non-renewable
monomer is used as a starting material while in the second
case, a renewable, non-toxic monomer is used. In other
words, this example highlights how one should use caution
and not consider each green chemistry principle in
isolation.
Among the different step-growth polymerization techniques, it is the interfacial type that is referred to as the most
atom economical. Another advantage of this method is that
higher molecular weight polymers are usually produced.[152,153] Interfacial polymerization is characterized
by the formation of polymers at or near a phase boundary of
two immiscible monomer solutions. These reactions are
carried out at lower temperatures, which results in lower
relative rates of side reactions. Monomer purity, in this case,
is not as critical as in typical step-growth polymerizations.
Nevertheless, interfacial polymerization has not been
applied widely industrially due to the high cost of the
reactive monomers and the significant solvent removal and
recovery step.[135] However, in the case of many other stepgrowth polymerizations, monomer purity is of prime
significance as it determines the possibility of formation
of unwanted by-products through side reactions; hence
higher atom economy is to be expected only when high
purity monomers are employed.
Contrary to step growth polymerizations, chain growth
polymerizations are highly atom efficient and all atoms in
the monomers are consumed during the course of
polymerization. Nonetheless, it is worth mentioning that
a chain growth polymerization might indeed lose atoms via
side reactions and waste generation. As discussed in Section
1, the generation of waste will be influenced by monomer
purity and control of the reaction conditions (e.g.,
temperature, ingredient feed rates).
In conclusion, atom economy (or E-factor) is an important
criterion to be accounted for, when considering sustainable
processes. Several methods to reduce waste formation in,
e.g., the fine chemical industry have been implemented and
resulted in higher atom economy. Polymer processes, for
the most part, tend to have an inherently high atom
economy due to the reaction mechanism. Achieving an
even higher atom economy may involve a trade-off among
several other equally or more important factors that

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Applying the Principles of Green Chemistry to Polymer . . .


www.mre-journal.de

contribute toward the attainment of a more sustainable


process. Often, one would need to employ an alternative
polymerization mechanism to obtain the same polymer
and this may not always be feasible.
1.8. Use safer solvents and reaction conditions: avoid
using solvents, separation agents, or other auxiliary
chemicals. If these chemicals are necessary, use
innocuous chemicals. If a solvent is necessary, water
is a good medium as well as certain eco-friendly
solvents that do not contribute to smog formation or
destroy the ozone
Solution polymerization continues to be an important
method to produce polymers with valuable properties. The
use of a solvent as a polymerization media prevents
increases in viscosity due to the generation of high
molecular weight polymer chains by diluting the reaction
mixture. This translates into improved heat transfer and
prevention of thermal runaway by absorbing the heat of
polymerization. Many polymer products have been developed over time, which present desirable molecular weight
qualities and other final properties that cannot be easily
met by other polymerization technologies (e.g., emulsion
polymerization). However, with strict regulation and
concerns over VOCs in various polymer production
processes, alternatives to conventional solvents are being
sought. Of course, alternatives may not be necessary if one
is working with water-soluble polymerization systems.
However, it is worth considering whether hazardous
wastewaters are produced in such processes. Some
important environmentally-friendly alternative solvents
are discussed below.
Monoterpenes (MTs) are a class of terpenes and have the
molecular formula C10H16. Terpenes are a broad class of
molecules found in nature and their structure may be cyclic
or acyclic with functional groups such as alcohols,
aldehydes, ketones, esters, and carboxylic acids. MTs have
several similarities with petroleum solvents based on
available data for dielectric constants and densities, and
they resemble aromatic solvents, such as toluene and
xylene.[154] Therefore, they have the potential to act as
solvents for a broad range of monomers and catalysts and
have the ability to replace non-renewable solvents in some
polymerization processes. The substitution of petroleumbased solvents such as toluene and methylene chloride
with MTs has been reported for ring-opening metathesis
polymerization (ROMP),[155] and for the synthesis of a wide
variety of polyesters, polyethers, polyamides, and unsaturated polyolefins.[156158] MTs have also been utilized for the
production of hyperbranched polymers in a one-pot process
that provides an economic alternative to multi-step
dendrimer synthesis.[159] The prepolymers produced using
MTs show good physical properties compared to the

www.MaterialsViews.com

network produced in the presence of toluene. However,


in the presence of MTs, chain transfer reactions were
observed which limit the formation of growing
insoluble networks; which can be important in some
processes.[154]
Ionic liquids (ILs) are organic salts composed of ions in
liquid form, close to room temperature.[160] In the last
decade, ILs have been considered the green media of the
future[161163] and have shown potential as non-volatile
organic solvents for polymerizations because of their nearzero vapor pressure, non-flammability, and ease of
production.[164] The term IL also includes neoteric
solvents, ionic fluids, liquid organic salts, and molten
salts.[165168] ILs are relatively non-volatile, therefore they
do not produce VOCs.[169172] They show good thermal
stability over a wide temperature range, and can thus be
utilized at higher temperatures (e.g., 800 8C).[169,173175]
Being both polar and non-coordinating solvents, ILs exhibit
good solvent properties for a wide range of monomers in
different chemical processes.[172,174] In addition, they have
the ability to interact through hydrogen bonding, and
dipolar and electrostatic interactions.[173174] Their highly
ionic character improves the reaction rates in various
reactions such as microwave-assisted organic synthesis
and polymerizations.[176]
The use of ILs as a polymerization solvent has been
reported for polycondensations, free-radical polymerizations, and ionic (anionic and cationic) polymerizations,
where they were mostly used in the synthesis of
polyamides, polyimides, and polyesters.[177,178] In some
cases, it was observed that while ILs were used as solvents
they may act as catalysts as well; e.g., in the polycondensation of phenol and formaldehyde.[179] Another major
advantage of using ILs is the absence of enzyme deactivation, which usually takes place in most polar solvents (e.g.,
methanol).[180] Free-radical polymerizations were also
performed in ILs and interestingly, the molecular weights
and rates of reactions were reported to be higher compared
to that in organic solvents.[174176,181,182] The rate constants
of propagation tend to be higher and the rates of termination
lower than in bulk or conventional solution polymerization.
This phenomenon has special significance for CRP, especially
atom transfer radical polymerization.[183]
Despite numerous advantages in using ILs, they exhibit
some technical challenges such as high viscosity. In
addition, they are sometimes sensitive to moisture,
which is not typically desirable in many chemical
syntheses. Furthermore, their toxicity has not been fully
studied.[21,184]
Supercritical fluids like supercritical carbon dioxide are a
small class of solvents which have been employed in
polymerization reactions. Supercritical CO2 has been
extensively studied for polymer production,[185] has
interesting physical properties, and poses few health

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

17

, S. Salehpour
M. A. Dube
www.mre-journal.de

hazards.[186] Supercritical fluids can have gas-like diffusivities which are very beneficial for reactions and have
liquid-like densities which allow solvation of many
chemicals.[185] In addition, they show changes in solvent
density with small changes in temperature or pressure. Of
course, CO2 is widely available and is usually released in
very large quantities as a by-product of other processes. In
general, using supercritical CO2 can be beneficial because
of the fact that it is inexpensive and non-toxic.[185]
However, there are certain constraints for its use as a
solvent in polymer synthesis. Contrary to its outstanding
solvent properties for small molecules, it is a poor solvent
for most high molecular weight polymers under mild
conditions (i.e., <100 8C and/or <350 bar).[187] For instance,
poly(methyl acrylate) requires 2 000 bar pressure at 100 8C
for a 105 g  mol1 polymer to dissolve in CO2. These
operating conditions are not cost effective. The only
polymers that exhibit considerable solubility in CO2
under mild operating conditions are non-crystalline
fluoropolymers and silicones.[188193]
Recently, fatty acid methyl esters (FAME or biodiesel)
produced from canola oil has been used as a polymerization
solvent, where solution polymerizations of four commercially important monomers [i.e., MMA, styrene, butyl
acrylate (BA), and VAc] were studied at 60 and
120 8C.[9,194] In the 1960 s, the use of methyl oleate, a major
component of biodiesel, as a polymerization solvent was
explored.[195] However, the high cost of the pure compound
prevented its widespread application. The solvating ability
of biodiesel has not been widely studied.[196] Biodiesel is
more commonly used as an alternative to petroleum diesel
for use in combustion engines. It is a fatty acid alkyl ester
(FAAE) produced via catalytic transesterification of vegetable oils, animal fats, or grease with an alcohol.[197] Biodiesel
fulfills the requirements of a good polymerization solvent;
it is environmentally benign and has low volatility, low
viscosity, and good solubility. In addition, its elevated
boiling point (e.g., 326 8C for canola-based biodiesel) points
to reduced workplace hazards; reactions can be carried out
at elevated temperatures without excessive pressure
buildup, aside from the contributions of the monomers.[9]
There has been mounting interest in carrying out polymerizations at elevated temperatures.[198] Advantages of using
high temperatures include decreasing the required concentration of chain transfer agents and initiators, and
increasing the reaction rate. However, there are disadvantages inherent to running reactions at elevated temperatures such as increased energy consumption, safety
considerations, and possible undesired side reactions that
may occur, i.e., intramolecular chain transfer and depropagation.[199] One drawback of using biodiesel as a highboiling solvent in solution polymerizations is the difficulty
in solvent removal from the final products. In contrast, for
polymerizations in traditional low-boiling solvents, the

18

solvent can be easily removed from the polymer product by


an evaporation process.
Nowadays, most chemical process industries try to avoid
using harmful solvents, separation agents, or other
auxiliary chemicals to achieve more environmentally
friendly processes. This is also of great concern to the
polymer industry and presents an area of high interest.
1.9. Increase energy efficiency: run chemical
reactions at ambient temperature and pressure
whenever possible
Polymerization reactions often proceed under conditions of
mass self-heating because of the exothermic nature of most
polymeric reactions.[200] The non-isothermal nature of
polymerization reactions has long been considered undesirable as it can lead to inconsistent polymer properties and
involves challenges in controlling the reaction temperature. However, in recent years, heat generation has been
considered a favorable aspect to be exploited in polymerization, especially in light of efforts to minimize the amount
of energy needed to run the reaction. These polymerization
reactions can be made to proceed in an almost selfpropagating mode and furthermore, the heat generated by
the polymerization could be used to preheat various input
streams. In other words, there is a case to be made for the use
of adiabatic polymerization conditions.[201203]
To explain the effects of temperature on chain growth
polymerization, e.g., it is useful to consider the general
polymerization rate equation:[135]:
Rp

kp
k0:5
t



2kd f I0:5 M

where [I] is the concentration of initiator or catalyst, [M] the


total monomer concentration, f the initiator efficiency,
and kp, kt, and kd are the propagation, termination, and
decomposition rate parameters
According to the polymerization rate equation (see
Equation 1), the effect of an increase in temperature can be
estimated by the change in the quantity kp(kd/kt)0.5. Based on
the fact that the kinetic rate parameters follow the Arrhenius
equation, the activation energy of the polymerization, ER,
which is a function of activation energy for propagation, EP,
activation energy for initiator decomposition, Ed, and
activation energy for bimolecular termination, Et, tends to
be high at elevated temperatures largely due to Ed.[135] This
indicates that the rate of polymerization increases strongly
with temperature. In bulk or solution chain growth
polymerization, the average polymer chain length decreases
significantly with increasing temperature. This occurs when
the initiation and bimolecular termination steps dominate
over other molecular mass development steps. When unimolecular termination (chain transfer to small molecules)

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Applying the Principles of Green Chemistry to Polymer . . .


www.mre-journal.de

dominates the molecular mass development, EP is found to be


less than Ef (activation energy for chain transfer reaction),
implying that the molecular mass can decrease with an
increase in temperature. Adjusting the chain transfer agents
or using a mixture of initiators with different kd values can
assist with maintaining a constant concentration of free
radicals, which results in a constant polymerization rate. This
ultimately would lead to a relatively uniform molecular
weight distribution.[204] Thus, in the case of an adiabatic
polymerization, one could potentially manipulate initiator
and chain transfer agent concentrations to compensate for
the effect of increasing temperature and therefore control the
polymer properties.
Photopolymerization reactions have major capabilities
compared to more traditional polymerizations and have
found application in many conventional products for over
three decades.[205] One of their most outstanding characteristics is the ability of photopolymers to cure rapidly at
ambient conditions.[206,207] Because of the low reaction
temperature, fewer side reactions (e.g., chain transfer) are
likely. In addition, the polymerization of monomers with
low ceiling temperatures (e.g., a-methylstyrene) or polymerization in the presence of temperature-sensitive
situations (e.g., in biomedical and dental applications)
are made possible. Photopolymerization has been used in a
wide range of applications including dentistry,[208210]
contact and other lenses,[211] coatings,[212,213] photolithography,[214,215] tissue engineering matrices,[216,217] and 3D
prototyping.[218,219] However, there are problems related to
volume shrinkage and stress, oxygen inhibition, and the
presence of unreacted monomer which can limit the
utilization of this method.[205,220222] The issue arises from
the fact that, in most photopolymerization reactions,
polymer formation is associated with a dramatic material
property change;[223] i.e., an initial liquid mixture with low
viscosity rapidly gets converted into a glassy polymer.
Several recent photopolymerization examples indicate that
this is an area of growth.[224228]
Another area of potential energy efficiency lies in ambient
temperature reactions such as ambient reversible addition
fragmentation chain transfer radical (RAFT) polymerization,
which is a highly versatile form of CRP.[229232] This
technique has been investigated for the polymerization
of many monomers that are able to polymerize under
conventional free radical polymerization conditions, providing outstanding control over their molecular weights; this
results in the synthesis of polymers with complex architectures.[233237] One should note, however, that the presence
of dithioester groups in the RAFT agents often lead to
polymers with undesirable color and odor.[238]
Microwave assisted polymerization (MAP) offers many
interesting advantages that begin with the possibility of
safe and easy polymer synthesis at elevated temperatures
(e.g., 200300 8C). It follows that one can expect shorter

www.MaterialsViews.com

reaction times, higher monomer conversions, more efficient (i.e., even) heating, and straightforward scale-up using
MAP.[239] Interesting examples in emulsion polymerization
have been reported for styrene,[240] MMA,[241,242] butyl
methacrylate,[243] and many others.[244] Examples over a
broader spectrum including synthesis, crosslinking, and
processing have also been reviewed.[245,246]
Microwave heating is perceived to be highly energy
efficient though there has been conflicting evidence
regarding this efficiency. A review of several published
works drew the conclusion that when considering the
energy profile of an entire process, greater overall energy
efficiency can be achieved through microwave heating in
comparison to conventionally heated reactions.[247] This is
primarily due to the ability to shorten reaction times by
conducting reactions at elevated temperatures and also
by considering the higher reaction yields and reduced
energy needs for separation equipment and mixing. It
should be noted that these efficiencies were realized
primarily at the larger reactor scale.
The search for energy efficient and cost-effective
polymerization techniques has become more critical of
late due to escalating energy costs. Most of the techniques
used for increasing the energy efficiency of polymerization
reactions are either in terms of running polymerizations
adiabatically or implementing photopolymerization, in
which UV rays are used to initiate the reaction. There is
ample room to exploit adiabatic polymerizations, particularly from the point of view of polymer property control,
though it is a mature technology. Photopolymerization also
appears to offer fertile ground for greater development.
1.10. Design chemicals and products to degrade after
use: design chemical products to break down into
innocuous substances after use so that they do not
accumulate in the environment
Polymers are employed in a multitude of applications, in
some cases, due to their lower weight, flexibility, durability,
and lifespan, ease of processing and lower cost. In addition,
the physical and mechanical properties of polymeric
materials often can be easily manipulated. As a result,
many polymeric materials have replaced traditional paper-,
metal-, and wood-based materials. Despite their many
advantages, most polymeric materials have extremely long
persistence times which often go well beyond the intended
practical lifespan of the material (e.g., packaging materials,
disposable cutlery, and cups). This can result either in
limiting the lifetime of landfill sites or, in the case of
incineration, the release of greenhouse gases (or worse) into
the environment. To further illustrate, polyolefins such as
PE and poly(propylene) are produced at a rate of 200 000 000
metric tonnes annually with PE alone being produced in
amounts exceeding 80 000 000 metric tonnes per year.[248]

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

19

, S. Salehpour
M. A. Dube
www.mre-journal.de

Since these polymers remain in the environment for years


without any significant decomposition of the polymer,
there is an illustrated need for the production of degradable
polymers.
It is worth noting some common confusion in terminology: biodegradable polymers are not necessarily biopolymers, which are made from renewable raw materials.
Biodegradability does not imply anything about the raw
materials used to produce the polymers. Instead, biodegradability refers to the polymers ability to breakdown into
energy and more elementary components such as carbon
dioxide, methane, water through the actions of microorganisms such as yeasts, bacteria, and algae within a
reasonable period of time.[249,250] Biodegradation usually
takes place via two pathways, aerobically, in which organic
matter is converted into CO2, energy, water, etc., by
microorganisms in the presence of oxygen or anaerobically,
in which organic matter is metabolized by microorganisms
in the absence of oxygen.[251] Thus, biodegradability
depends on the chemical composition of the polymer and
biodegradable polymers can be made from either renewable raw materials or fossil-fuel based feedstocks.
A second important issue revolves around the mechanism of polymer degradation. One may mechanically
degrade a polymer whereas biodegradation is a certified
performance characteristic as can be found in ISO 17088, EN
13432 standards in Europe and the ASTM D 6400 standard
in North America. For example, in order to be compliant
with EN 13432, a polymer must be converted to carbon
dioxide by over 90% within 180 d under certain conditions
of humidity, temperature, and oxygen level.[252]
Biodegradable polymers can be obtained from both
renewable as well as non-renewable raw materials. Major
non-renewable and fossil-based monomers leading to
biodegradable polymers are: butanediol, dicarboxylic acids,
adipic acid, terephthalic acid, and succinic acid.[252] Among
the bio-sourced biodegradable polymers, PLA is a commercial polymer derived from lactic acid which is also highly
biocompatible.[5760,250254] PLA which can be produced
from sugar and starch over a wide range of molecular
weights has been recently used in various applications such
as packaging and rigid thermoforms.[61,62] This renewable
polymer possesses chemical and physical properties which
make it a suitable replacement for widely used polymers
such as PET and PS in the packaging industry.[255]
Although PLA has many interesting properties, it has some
drawbacks which limit its application. For example, compared to PS or PET it does not show good mechanical strength
and barrier properties at higher temperatures.[61,82,255] Many
researchers have been focusing on resolving these problems
by different methods including, most notably, copolymerizations, reactive extrusion, and blending.[256,257]
Another class of biodegradable polymers are polyhydroxyalkanoates (PHAs). These polymers, which have

20

structural similarities to PLA, are also produced from


renewable feedstocks like sugar and starch and polyesters
which differ in chain lengths and in the position of their
hydroxyl groups.[258] This class of polymers with different
chemical structures has a wide range of chemical and
physical properties and has applications in biomedical and
non-biomedical industries such as in biodegradable
films.[259261] However, polyesters are prone to chemical
hydrolysis, and in many cases, suitable co-monomers
should be involved in their syntheses in order to achieve
desirable mechanical properties and shelf lives.[262]
Another class of widely available, renewable, biodegradable, and inexpensive starting materials is starch. Starch is
found as discrete particles (starch granules) in plants. Pure
starch is a very water and temperature sensitive polysaccharide compound and has very limited applications in the
plastic and fiber industries. In order to modify the
hydrophobicity of starch and improve its stability against
solvents, cross-linking of polymer chains and partial
substitution of hydroxyl groups has been studied.[263]
Nonetheless, these methods can increase the cost of
production and lead to a non-biodegradable product.[252]
Other polymers that are designed to be biodegradable
include poly(aspartate) (synthesized by the polycondensation reaction of aspartic acid via poly(succinimide),
followed by hydrolysis), and poly(ethylene glycol) (a water
soluble and completely biodegradable polymer at lower
molecular weights).[264268]
Currently, biodegradable polymers are largely regarded
as specialty plastics for selected applications where
biodegradability adds value. Extensive research continues
to be performed on the development of many more
biodegradable polymers such as poly(aspartate), poly
(ethylene glycol), and PVA.[269,270]
1.11. Analyze in real time to prevent pollution:
include in-process real-time monitoring and control
during syntheses to minimize or eliminate the
formation of by-products
Real-time polymerization monitoring techniques are important for the synthesis of polymers with pre-specified
properties and for the prevention of waste. These methods
improve and/or enable process control procedures and
prevent accidents such as runaway reactions, which
ultimately serves to protect operators and equipment
(see Section 12). Most real-time methods are costeffective, non-destructive and have little if any negative
effect on the environment. During the last two decades,
there has been significant growth in sensor technologies
and methods for monitoring polymerizations.[271273]
The perfect reaction monitoring tool should be cost
effective, easy-to-install, explosion-proof, and preferably
calibration and maintenance free. However, until now, only

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Applying the Principles of Green Chemistry to Polymer . . .


www.mre-journal.de

pressure and temperature sensors have come close to


meeting these ideal conditions.[273] Nevertheless, pressure
and temperature sensors only measure the state of the
reactor and do not directly monitor important, ongoing
reaction changes like composition of the reaction
mixture. Some common monitoring techniques are described below:
1.11.1. Reaction Calorimetry
Most chemical reactions are either exothermic or endothermic, in polymerization usually conversion of monomer
to polymer generates heat which can be measured to
monitor polymerization.[201,202,274276] One well-known
method for controlling and monitoring polymerizations
is heat balance calorimetry. In this method, which has been
used industrially for some time, the energy balance of the
jacket cooling medium is measured.[277] However, several
factors can affect the heat balance calculation, which can
compromise the results, such as heat loss by radiation,
sampling from the reactor and heat dissipation as a result of
mixing in reaction mediums with high viscosity. In
addition to the above possible sources of error in the
calorimetric method, this technique has shortcomings
when it comes to systems with several monomers and
variables; this is the case in copolymerization systems.
Therefore, supplementary analyses and data collection
about specific monomers should be obtained via other
methods for proper reaction monitoring.

they usually do not require any additional chemicals or


sample pre-processing (e.g., dilution). They are also nondestructive and prevent waste formation. Therefore, these
robust techniques enable the fast real-time analysis of
samples without the need to extract the samples from the
reaction vessel. One concern about monitoring polymerizations with optical sensors is the formation of a polymer
film on the optic probe[281] which might be resolved by
cleaning the sensor with by-passes, process feeds (e.g.,
solvents and monomers), or extraction fixtures.[273] Monitoring high pressure polymerization systems is also a
significant challenge. Reactors are often cleaned by means
of high-pressure jets of cleaning solutions and a direct hit of
a sensor probe by this jet might lead to permanent damage
to the sensor.[273] Brief descriptions of some common
spectroscopic techniques are presented below. More details
about on-line sensors can be found in a recent review
paper.[271]
1.11.4. Infrared (IR) Spectroscopy

This method is the most widely used on-line monitoring


technique for chemical reactions in the petrochemical
industry[278] and for organic syntheses[279] due to its wealth
of qualitative and quantitative information.[280] The sample
from the reaction mixture should be fed into the GC
instrument for analysis via a circulation loop or a transfer
line. However, there is usually a time-lag between the
sampling time and the analytical results. Therefore, in cases
with long delays, this method can be considered an off-line
monitoring technique rather than a real-time one. Another
drawback of this method for monitoring polymerizations is
that the highly viscous or solid polymer product can clog
transfer lines, valves, and columns and may require additional
dilution. Nonetheless, GC is a common method for monitoring
residual monomer and VOCs in polymerization mixtures.[273]

IR spectroscopy is a popular analytical method for


preparative and analytical chemistry which provides
structural and kinetic information in a non-destructive
and waste-free way.[282] The IR band is divided into three
regions: near-, mid-, and far-IR. The near-infrared (NIR)
spectrum extends from 13 000 to 4 000 cm1 and it provides
information on overtones or a combination of the
fundamental stretching bands occurring from 3 000 to 1
700 cm1. On the other hand, the mid-infrared (MIR)
spectrum extending from 4 000 to 400 cm1 gives
information on fundamental molecular vibrations and
usually is the preferred choice owing to the unmatched
wealth of molecular information contained in this portion of
the electromagnetic spectrum.[274] In situ Fourier transform
infrared (FTIR) spectroscopy which simultaneously collects
spectral data in a wide spectral range, is a modern
polymerization monitoring technique that is well suited
for obtaining real-time structural and kinetic information.
The rapid-scanning capability of FTIR spectroscopy enables
monitoring of the time-dependent intensity changes of
absorbance of the molecules throughout the polymerization.[283] In order to monitor water-rich reactions, internal
reflection spectroscopy (often called attenuated total
reflectance or ATR) can be used. Real-time monitoring of
conversion of reaction and composition are possible using
ATR-FTIR spectro- scopy.[271,281,284,285]

1.11.3. Optical Spectroscopy

1.11.5. Raman spectroscopy

The composition of the reaction mixture and the progress of


polymerizations can be monitored using different optical
methods via in-line or fast off-line analysis. Spectroscopic
techniques have many advantages over some other
monitoring techniques such as wet chemistry or GC as

Similar to IR spectroscopy, the molecular structure, and


properties of reaction components can be analyzed using
Raman spectroscopy based on their vibrational transitions.
However, most Raman monitoring techniques use fiberoptic probes[286] which employ a single frequency of

1.11.2. Gas Chromatography (GC)

www.MaterialsViews.com

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

21

, S. Salehpour
M. A. Dube
www.mre-journal.de

radiation to irradiate the sample, in contrast to IR


spectroscopy which uses a range of frequencies.[287] This
method is suitable for monitoring high water content
reactions such as emulsion and suspension polymerizations and is very suitable for monitoring the signal
from carboncarbon double and triple bonds.[273] It is
worth mentioning that Raman spectroscopy can be
strongly affected by properties of the medium such as
turbidity.[288]
1.11.6. UV/Vis Spectroscopy
UV spectroscopy is a well-known analytical technique
which is quite fast and requires only a small amount of
sample. Therefore, it is a suitable method for on-line
monitoring. The UVVis part of the spectrum, extending
from 200 to 800 nm, corresponds to the excitation of the
outer electrons of a molecule.[271] However, only vinyl
monomers show adsorption in this spectral range.[289]
Successful monitoring of MMA and styrene polymerization
was performed via this technique.[273]
1.11.7. Ultrasound
Ultrasonic methods are cost-effective, non-invasive realtime monitoring techniques. Certain properties of a medium
such as velocity, viscosity, and elasticity can be measured
utilizing ultrasonic waves.[290] Using semi-empirical models, properties like sound velocity can be related to the
conversion of polymerization.[291] However, this technique
has certain disadvantages, whereas acoustic properties of
the material and accurate models are required in order to
perform reliable data analysis.[292]
Real time monitoring of reactions is expected to improve
significantly by implementation of new technologies in the
future and there are several advanced techniques such as
acoustic emission, online NMR spectroscopy, ion mobility
spectroscopy, and process tomography that are under
study[273] and can improve online monitoring of polymerization and prevent the formation of by-products and
pollution.
1.12. Minimize the potential for accidents: design
chemicals, and their forms (solid, liquid, or gas) to
minimize the potential for chemical accidents
including explosions, fire, and release to the
environment
The potential for accidents in the polymer industry is large
due to the variety and volume of materials used as well as
the presence of these materials in different states (i.e., gas,
liquid, or solid). As discussed earlier, many polymerizations
are conducted under fairly aggressive conditions (i.e.,
elevated temperatures and pressures) and many are quite
exothermic, posing the threat of thermal runaway. The
potential for accidents can be considered from two

22

perspectives: (i) during the synthesis and processing stages


of polymer production and (ii) accidents occurring when
using the final polymer product. The latter relates primarily
to the release of hazardous materials in the event of a fire. It
is often mitigated by the use of fire retardants and oxygen
scavengers which serve to prevent degradation and release
of harmful materials to the environment.[293295] The use of
environmentally benign additives was discussed earlier
in Section 2. The focus of the present discussion is on
minimizing incidents during polymer synthesis and
processing.
A principal reason for accidents in polymer synthesis
relates to the transport and handling of chemicals. These
chemicals, particularly monomers and initiators, are fairly
reactive and usually have the potential to get triggered by
external factors, and cause damage to the workplace and
harm to workers. One significant incident occurred in the
Scottish seaport of Grangemouth, in 2006 where 24 000 L of
divinyl benzene (DVB) contained in a tanker exposed to
sunlight on the docks self-polymerized.[296] DVB is commonly used as a cross-linker for the production of ion
exchange resins. After the accident, firefighters sealed off
the area within a 500 m range and residents were forced to
remain in their homes for 24 h because of the irritating
characteristics of DVB towards skin and eyes. The resulting
vapor cloud was eventually completely dispersed by the
wind.[296]
Conditions in handling and storage of monomers, most
notably temperature, are critical factors that influence
the self-polymerization of monomers during transport.
In most cases, inhibitors are added to the monomer. For
example, 4-tert-butylcatechol (TBC) is added at a level of
9001 200 ppm by weight to inhibit the self-initiated
polymerization of DVB. However, it was noted that the
addition of TBC alone was not sufficient to stabilize DVB in
the Grangemouth incident and temperature and the level
of oxygen in the tank also were very important factors.
Obviously, each particular scenario poses its own hazards
but temperature and the presence of radical scavengers do
play important roles.
In addition to handling and storage, it should be noted
that most polymerizations are exothermic reactions.
Whenever cooling fails, uncontrolled runaway reaction
may take place, leading to increased temperatures and a
potential chain of events leading to disaster. As a result of
reaction runaway, increase in vapor pressures and side
reactions may occur. In these cases, appropriate process
control techniques should be implemented.[293] This could
include manipulating not only the heat transfer equipment
but the inputs of the various reaction ingredients, and in
particular the initiator. Several process control methods and
equipment and safety devices exist to control the polymerizations such as the simple use of pressure relief valves or
emergency cooling systems.

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Applying the Principles of Green Chemistry to Polymer . . .


www.mre-journal.de

Table 2. Risk Assessment for PTT, PHA, PET, and PE.[297]

Product

Risk (YOLL/Ton product)

Poly(trimethylene terephthalate) (PTT)

Polyhydroxyalkanoates (PHA)

From ethylene oxide

0.003821

From acrolein

0.0043031

Via anaerobic fermentation of dextrose

0.003001

Via anaerobic fermentation of glycerol

0.003227

Via aerobic fermentation

0.003039

Via fermentation

0.002501

Poly(ethylene terephthalate) (PET)

0.003211

Poly(ethylene) (PE)

0.004949

Other considerations include: the implementation of


polymerization techniques where the medium acts as a
heat sink (e.g., solution, emulsion); fouling, which can in the
short and long term, change the heat transfer characteristics of the reactor; mixing which plays a role both in heat
transfer and fouling; and regular equipment maintenance
and inspection.
In the end, it is important to consider the associated
risks in the synthesis, processing, and use of the polymer
products from a cradle to grave perspective. A study
combining classical risk assessment methods with statistics on technological disasters, accidents, and work-related
illnesses was carried out to assess the associated risks of
certain polymers, such as poly(trimethylene terephthalate)
(PTT), PHA, PET, and PE.[297] The total risk to human health
throughout a chemicals life cycle was estimated by
summing up the following risks:[297]
(1) External risks due to the release of emissions from
regular operation;
(2) External risks due to technological disasters;
(3) Risks of work-related accidents;
(4) Risks of work-related illnesses.
Using this risk assessment strategy, estimates for the total
number of years of life lost (YOLL) per unit of product
throughout the process chain were calculated as can be seen
in Table 1 for PTT, PHA, PET, and PE.[297] For PTT synthesis,
given the possibility of both petrochemical and bio-based
production routes, it is possible to compare these routes for
their risk to human health. On the other hand, PET and PE are
fossil-based polymers and can be compared to the alternative
PHA which is produced from bio-based feedstocks. Based on
risk assessment results presented in Table 2, the risks to
human health of bio-based polymers are significantly lower
than those of the petrochemical polymers.
Accidents related to the polymerization process are
largely due to the handling of several raw materials used in
the process such as monomers, solvents, and initiators as
well as the highly exothermic nature of the polymer-

www.MaterialsViews.com

izations. These issues can be mostly resolved by implementing proper environmental and advanced process
controls. In addition, substitution of petrochemical feedstocks with bio-based feedstocks may also contribute to
decreasing the potential for accidents.

2. Conclusion
As an economically important and technologically vast and
interesting field, polymer reaction engineering is a prime
target for transformation towards greater sustainability.
There is not a single, unique approach for the transformation of polymerization processes into more sustainable
ones, but rather a coordination of several steps can result in
an environmentally friendly polymerization process. This
transformation is being made possible via the 12 principles
of green chemistry:[1]
(1)
(2)
(3)
(4)
(5)
(6)
(7)
(8)
(9)
(10)
(11)
(12)

Prevent waste;
Design safer chemicals and products;
Design less hazardous chemical syntheses;
Use renewable feedstock;
Use catalysts not stoichiometric reagents;
Avoid chemical derivatives;
Maximize atom economy;
Use safer solvents and reaction conditions;
Increase energy efficiency;
Design for degradation after use;
Analyze in real time to prevent pollution; and
Minimize the potential for accidents.

As noted, while there is always room for improvement,


several of the principles are essentially being achieved, for
the most part, in modern-day polymer production; these
include principles 5, 6, 7, 11, and 12. Principles 1, 9, and 10
can be considered as being well underway but still showing
promise for important developments. For example, in terms
of principle 9, exploiting photopolymerization may require
further advances in catalysis while the use of adiabatic

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

23

, S. Salehpour
M. A. Dube
www.mre-journal.de

polymerization techniques implies a need for improved


polymer property modeling, monitoring, and control.
Another example, in terms of principle 10, lies in the
ongoing development of biodegradable polymers.
One can argue that those principles seeming to have the
greatest potential for significant environmental benefit are
principles 2, 3, 4, and 8. Principle 2, concerns the design of
safer chemicals and products. The replacement of toxic
additives with non-toxic alternatives and the reduction in
VOCs (e.g., from solvents) in polymer products appear to
offer many challenges and significant room for improvement. Designing less hazardous chemical syntheses (Principle 3) offers enormous opportunities from the replacement of solvents to the use of less toxic monomers
in polymer production. The use of renewable feedstocks
(Principle 4) is a highly active area of research,
despite limitations to the supply of these feedstocks.
Achieving Principle 8, the use of safer solvents and reaction
conditions, also appears to have elicited significant research
effort.
Ultimately, keeping an eye towards all 12 green
chemistry principles should become standard practice for
all polymer scientists and engineers. In this way, the
inevitable and necessary transformation of polymer
production towards a more sustainable future will be
greatly facilitated.
Acknowledgments: The authors gratefully acknowledge the
support of the Natural Sciences and Engineering Research Council
(NSERC) of Canada. Abebe Essayas and Zahra Dastjerdi are also
acknowledged for their assistance in the collection of literature
data for this paper.

Received: February 23, 2013; Revised: July 5, 2013; Published


online: August 29, 2013; DOI: 10.1002/mren.201300103
Keywords: green chemistry; polymerization; polymer reaction
engineering; renewable resources; sustainability

[1] P. T. Anastas, J. C. Warner, Green Chemistry: Theory and


Practice, Oxford University Press, New York 2000, p. 30.
[2] P. H. H. Araujo, C. Sayer, J. G. R. Poco, R. Giudici, Polym. Eng.
Sci. 2002, 42, 1442.
[3] M. Aerts, J. Meuldijk, M. Kemmere, J. Keurentjes, Macromol.
Symp. 2011, 302, 297.
[4] US Patent 5 767 231 1998, Agomer Gesellschaft Mit
Beschrankter Haftung, invs.: D. Arnoldi, V. Schull.
[5] US Patent 5 376 703 1994, Hoechst Ag, invs.: H.U. Huth,
E. Noelken, H. Rinno.
[6] S. K. Ooi, S. Biggs, Ultrason. Sonochem. 2000, 7, 125.
[7] K. Kabiri, S. Hesarian, A. Jamshidi, M. J. Zohuriaan-Mehr,
H. Boohendi, M. R. Poorheravi, S. A. Hashemi, F. AhmadKhanbeigi, J. Appl. Polym. Sci. 2011, 120, 2716.
[8] S. Copelli, M. Derudi, J. Sempere, E. Serra, A. Lunghi, C. Pasturenzi,
R. Rota, J. Hazard. Mater. 2011, 192, 8.

24

[9] S. Salehpour, M. A. Dube, Green Chem. 2008, 10, 329.


[10] G. E. Fonseca, T. F. McKenna, M. A. Dube, Chem. Eng. Sci. 2010,
65, 2797.
[11] A. Salvini, L. M. Saija, M. Lugli, G. Cipriani, C. Giannelli, J.
Adhes. Sci. Technol. 2010, 24, 1629.
[12] M. A. Dube, A. Penlidis, J. Polym. Sci., Polym. Chem. 1996, 35,
1659.
[13] G. V. Schulz, G. Henrici-Olive, S. Olive, Makromol. Chem.
1959, 31, 88.
[14] C. H. Bamford, G. C. Easmond, D. Whittle, Polymer 1969, 10,
771.
[15] M. Stickler, D. Panke, A. E. Hamielec, J. Polym. Sci., Polym.
Chem. 1984, 2243.
[16] L. P. Real, J. L. Gardette, A. Pereira Rocha, Polym. Degrad.
Stabil. 2005, 88, 357.
[17] M. Weiss, J. Haufe, M. Carus, M. Brando, S. Bringezu,
B. Hermann, M. K. Patel, J. Ind. Ecol. 2012, 16, S169.
[18] D. Sisodiya, P. Pandey, K. Dashora, J. Pharm. Res. 2012, 5, 852.
[19] Y. Huang, C. Wong, J. Zheng, H. Bouwman, R. Barra, B.
m, L. Neretin, M. Wong, Environ. Int. 2012, 42, 91.
Wahlstro
[20] J. Westerdahl, M. Belhaj, T. Rydberg, J. Munthe, R. Darbra,
A. a
gueda, S. Heise, L. Ziyang, in: Global Risk-Based Management of Chemical Additives I, Vol. 18, B. Bilitewski, R. M.
Darbra, D. Barcelo, Eds., Springer-Verlag, Berlin 2012, p. 57.
[21] T. Welton, Chem. Rev. 1999, 99, 2071.
[22] M. G. Benton, C. S. Brazel, Polym. Int. 2004, 53, 1113.
[23] Y. Y. Xie, Synth. Commun. 2005, 35, 1741.
[24] M. Xanthos, in: Functional Fillers for Plastics, 2nd edition
M. Xanthos, (Ed., Wiley-VCH Verlag GmbH, Weinheim 2010,
p. 3/18.
[25] F. P. La Mantia, M. Morreale, Compos. Part A: Appl. Sci. 2011,
42, 579.
[26] S. N. Monteiro, V. Calado, R. J. S. Rodriguez, F. M. Margem, J.
Mater. Res. Technol. 2012, 1, 117.
[27] U. Casado, N. E. Marcovich, M. I. Aranguren, M. A. Mosiewicki,
Polym. Eng. Sci. 2009, 49, 713.
[28] C. Alves, P. M. C. Ferrao, A. J. Silva, L. G. Reis, M. Freitas, L. B.
Rodrigues, D. E. Alves, J. Clean. Prod. 2010, 18, 313.
[29] S. M. Luz, P. M. C. Ferro, C. Alves, M. Freitas, A. Caldeira-Pires,
Mater. Sci. Forum 2010, 636637, 226.
[30] E. Hablot, R. Matadi, S. Ahzi, L. Averous, Compos. Sci. Technol.
2010, 70, 504.
[31] H. Deka, M. Misraa, A. Mohanty, Ind. Crop. Prod. 2013, 41, 94.
[32] X. Z. Tang, P. Kumar, S. Alavi, K. P. Sandeep, Crit. Rev. Food Sci.
2012, 52, 426.
[33] S. J. Eichhorn, A. Dufresne, M. Aranguren, N. E. Marcovich,
J. R. Capadona, S. J. Rowan, C. Weder, W. Thielemans,
M. Roman, S. Renneckar, W. Gindl, S. Veigel, J. Keckes,
H. Yano, K. Abe, M. Nogi, A. N. Nakagaito, A. Mangalam, J.
Simonsen, A. S. Benight, A. Bismarck, L. A. Berglund, T. Peijs, J.
Mater. Sci. 2010, 45, 1.
[34] D. F. Cadogan, C. J. Howick, in: Kirk-Othmer Encyclopedia of
Chemical Technology, Vol. 19, John Wiley and Sons, New
York 1996, p. 19/258.
[35] J. H. Kang, K. Kito, F. Kondo, J. Food Protect. 2003, 66,
1444.
[36] C. Brede, P. Fjeldal, I. Skjevrak, H. Herikstad, Food Addit.
Contam. 2003, 20, 684.
[37] J. D. Meeker, S. Sathyanarayana, S. H. Swan, Philos. Trans.
Roy. Soc. B 2009, 364, 2097.
[38] J. Wang, R. Che, W. Yang, J. Lei, Polym. Int. 2011, 60, 344.
[39] A. M. Nelson, T. E. Long, Polym. Int. 2012, 61, 1485.
[40] C. C. Geiger, J. D. Davies, W. H. Daly, J. Polym. Sci., Polym.
Chem. 1995, 33, 2317.

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Applying the Principles of Green Chemistry to Polymer . . .


www.mre-journal.de

[41] J. ukaszczyk, B. Janicki, M. KaczMRENk, Eur. Polym. J. 2011,


47, 1601.
[42] V. R. Dhara, R. Dhara, Arch. Environ. Health 2002, 57, 391.
[43] S. Chen, Q. Wang, T. Wang, J. Polym. Res. 2012, 19, 1.
[44] T. Keicher, W. Kuglstatter, S. Eisele, T. Wetzel, H. Krause,
Propellants Explos. Pyrotech. 2009, 34, 210.
[45] M. Helou, J.-F. Carpentier, S. M. Guillaume, Green Chem. 2011,
13, 266.
[46] Z. Li, Y. Zhao, S. Yan, X. Wang, M. Kang, J. Wang, H. Xiang,
Catal. Lett. 2008, 123, 246.
[47] M. G. A. Vieira, M. A. da Silva, L. O. dos Santos, M. M. Beppu,
Eur. Polym. J. 2011, 47, 254.
[48] A. K. Mohanty, A. Wibowo, M. Misra, T. Drzal, Polym. Eng. Sci.
2003, 43, 1151.
[49] T. Yorifuji, M. Noguchi, T. Tsuda, E. Suzuki, S. Takao, S.
Kashima, Y. Yanagisawa, J. Occup. Health 2012, 54, 79.
[50] E. J. North, R. U. Halden, Rev. Environ. Health 2013, 28, 1.
[51] M. Yoshioka, Biopolym. Online 2005, DOI: 10.1002/
3527600035.bpol9010, accessed: August 2013.
[52] S. Salehpour, M. A. Dube, Macromol. Chem. Phys. 2011, 212,
1284.
[53] S. Salehpour, C. J. Zuliani, M. A. Dube, Eur. J. Lip. Sci. Technol.
2012, 114, 92.
[54] M. Yoshioka, T. Miyazaki, N. Shiraishi, Mokuzai Gakkaishi
1996, 42, 406.
[55] N. Shiraishi, T. Matsunaga, T. Yokota, J. Appl. Polym. Sci. 1979,
24, 2361.
[56] M. Yoshioka, Y. Uehori, H. Toyosaki, T. Hashimoto, N.
Shiraishi, New Zealand FRI Bull. 1992, 176, 155.
[57] H. Tsuji, in: Biopolymers. Polyesters III. Applications and
Commercial Products, 1st edition, Y. Doi, A. Steinbuchel, Eds.,
Wiley - VCH Verlag GmbH, Weinheim 2002, p. 129.
[58] A. C. Albertson, I. K. Varma, Adv. Polym. Sci. 2002, 157, 1.
[59] K. E. Ulrich, S. M. Cannizzaro, R. S. Langer, K. M. Shakesheff,
Chem. Rev. 1999, 99, 3181.
[60] E. S. Lipinsky, R. G. Sinclair, Chem. Eng. Prog. 1986, 82, 26.
[61] R. Auras, B. Harte, S. Selke, Macromol. Biosci. 2004, 4, 835.
[62] P. R. Gruber, R. E. Drumright, D. E. Henton, Adv. Mater. 2000,
12, 1841.
[63] A. P. Po
go, B. Siebum, M. J. A. Van Luyn, X. J. G. Y. V. Seijen,
A. A. Poot, D. W. Grijpma, J. Feijen, Tissue Eng. 2003, 9,
981.
[64] J. J. Marler, J. Upton, R. Langer, J. P. Vacanti, Adv. Drug
Delivery Rev. 1998, 33, 165.
[65] A. P. Po
go, M. J. A. Van Luyn, L. A. Brouwer, P. B. van Wachem,
A. A. Poot, D. W. Grijpma, J. Feijen, J. Biomed. Mater. Res., A
2003, 67A, 1044.
[66] A. C. Albertson, M. Eklund, J. Polym. Sci. Polym. Chem. 1994,
32, 265.
[67] J. J. Verendel, T. L. Church, P. G. Andersson, Synthesis 2011, 11,
1649.
[68] D. P. Pfister, Y. Xia, R. C. Larock, ChemSusChem 2011, 4, 703.
[69] Y. Vandenberghe, Coat. Agenda Eur. 1998, 106.
[70] A. M. Fernandez, U. Held, A. Willing, W. H. Breuer, Prog. Org.
Coat. 2005, 53, 246.
[71] T. Benvegnu, J.-F. Sassi, Top. Curr. Chem. 2010, 294, 143.
[72] N. A. Smirnova, E. A. Safonova, Russ. J. Phys. Chem. A 2010, 84,
1695.
[73] C. Rondel, B. Portet, I. Alric, Z. Mouloungui, J. F. Blanco, F.
Silvestre, J. Surfactants Deterg. 2011, 14, 535.
[74] J. Davis, Eng. Plast. 1996, 9, 403.
[75] E. D. Weil, S. V. Levchik, M. Ravey, W. Zhu, Phosphorus Sulfur
1999, 144146, 17.
[76] S. Y. Lu, I. Hamerton, Prog. Polym. Sci. 2002, 27, 1661.

www.MaterialsViews.com

[77] R. J. Law, C. R. Allchin, J. de Boer, A. Covaci, D. Herzke, P.


Lepom, S. Morris, J. Tronczynski, C. A. de Wit, Chemosphere
2006, 64, 187.
[78] L. S. Birnbaum, D. F. Staskal, Environ. Health Persp. 2004, 112,
9.
[79] B. A. Howell, K. E. Carter, J. Therm. Anal. Calorim. 2010, 102,
493.
[80] A. Gandini, Macromolecules 2008, 41, 9491.
[81] G. W. Coates, M. A. Hillmyer, Macromolecules 2009, 42, 7987.
[82] C. K. Williams, M. A. Hillmyer, Polym. Rev. 2008, 48, 1.
[83] A. Gandini, in: Biocatalysis in Polymer Chemistry, K. Loos,
G. M. Guebitz, Eds., Wiley - VCH Verlag GmbH, Weinheim
2011, p. 1/33.
[84] J. O. Metzger, U. Bornscheuer, Appl. Microbiol. Biot. 2006, 71,
13.
[85] M. A. R. Meier, J. O. Metzger, U. S. Schubert, Chem. Soc. Rev.
2007, 36, 1788.
ner, Y. Yaci, A. Tuncer Erciyes, Prog. Polym. Sci.
[86] F. Seniha Gu
2006, 31, 633.
[87] L. Montero De Espinosa, M. A. R. Meier, Eur. Polym. J. 2011, 47,
837.
rk, S. H. Ku
sefolu, J. Appl. Polym. Sci. 2011, 121, 2976.
[88] C. ztu
, L. Montero de Espinosa, J. Carles Ronda, G. Lligadas,
[89] M. Galia
diz, Eur. J. Lipid Sci. Technol. 2010, 112, 87.
V. Ca
[90] M. Desroches, M. Escouvois, R. Auvergne, S. Caillol, B.
Boutevin, Polym. Rev. 2012, 52, 38.
[91] H. Sugimoto, S. Inoue, Kobunshi Ronbunshu 2005, 62, 131.
[92] S. Salehpour, M. A. Dube, J. Macromol. Sci., A 2012, 49, 103.
[93] R. T. Mathers, J. Polym. Sci. Polym. Chem. 2012, 50, 1.
[94] T. R. Carlson, T. P. Vispute, G. W. Huber, ChemSusChem 2008,
1, 397.
[95] A. Gandini, Polym. Chem. 2010, 1, 245.
[96] A. Brust, F. W. Lichtenthaler, Green Chem. 2013, 15, 1368.
[97] A. Gandini, T. M. Lacerda, A. J. F. Carvalho, Green Chem. 2013,
15, 1514.
[98] S. Bloembergen, I. McLennan, D. I. Lee, J. Van Leeuwen,
Paper3608 2008, 9, 46.
[99] C. P. Klass, Paper3608 2007, 1, 30.
[100] S. Bloembergen, I. J. McLennan, J. Van Leeuwen, D. I. Lee,
presented at 64th Appita Annual Conference and Exhibition,
Melbourne April, 2010.
[101] M.-A. Tehfe, J. Lalev, D. Gigmes, J. P. Fouassier, Macromolecules 2010, 43, 1364.
[102] M. Firdaus, L. Montero de Espinosa, M. A. R. Meier,
Macromolecules 2011, 44, 7253.
[103] C. Vilela, L. Cruciani, A. J. D. Silvestre, A. Gandini, RSC Adv.
2012, 2, 2966.
[104] S. Munoz-Guerra, High Perform. Polym. 2012, 24, 9.
[105] J.-M. Raquez, M. Deleglise, M.-F. Lacrampe, P. Krawczaka,
Prog. Polym. Sci. 2010, 35, 487.
, V. Ca
diz, React. Funct.
[106] J. C. Ronda, G. Lligadas, M. Galia
Polym. 2013, 73, 381.
[107] R. T. Mathers, M. A. R. Meier, Green Polymerization Methods:
Renewable Starting Materials, Catalysis and Waste Reduction, Wiley-VCH Verlag GmbH, Weinheim 2011.
[108] P. T. Anastas, M. M. Kirchhoff, T. C. Williamson, Appl. Catal.
A-Gen. 2001, 221, 3.
[109] T. Lester, in: Chemistry of Waste Minimization, J. H. Clark,
(Ed., Blackie Academic & Professional, London 1995, p. 1/16.
[110] G. Cecchin, G. Morini, F. Piemontesi, in: Kirk-Othmer
Encyclopedia of Chemical Technology Online, Vol. 19, John
Wiley and Sons, New York 2001, p. 502/554.
[111] D. S. Breslow, N. R. Newburg, J. Am. Chem. Soc. 1957, 79,
5072.

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

25

, S. Salehpour
M. A. Dube
www.mre-journal.de

[112] P. C. Moehring, N. J. Coville, J. Organomet. Chem. 1994, 1,


479.
[113] V. K. Gupta, S. Satish, I. S. Bhardwaj, J. Macromol. Sci. R. M. C.
1994, 34, 439.
[114] H. H. Brinzinger, D. Fischer, R. Muelhaupt, B. Rieger, R. M.
Waymouth, Angew. Chem., Int. Ed. 1995, 34, 1143.
[115] W. Kaminsky, M. Arndt, Adv. Polym. Sci. 1997, 127, 143.
[116] H. G. Alt, A. Koeppl, Chem. Rev. 2000, 100, 1205.
[117] G. W. Coates, Chem. Rev. 2000, 100, 1223.
[118] L. Resconi, L. Cavallo, A. Fait, F. Piemontesi, Chem. Rev. 2000,
100, 1253.
[119] S. D. Ittel, L. K. Johnson, M. Brookhart, Chem. Rev. 2000, 100,
1169.
[120] G. J. P. Britovsek, M. Bruce, V. C. Gibson, B. S. Kimberley, P. J.
Maddox, S. Mastroianni, S. J. McTavish, C. Redshaw, G. A.
mberg, A. J. P. White, D. J. Williams, J. Am. Chem.
Solan, S. Stro
Soc. 1999, 121, 8728.
[121] Q. Lou, D. A. Shipp, ChemPhysChem 2012, 13, 3257.
[122] A. Nese, Y. Li, S. S. Sheiko, K. Matyjaszewski, ACS Macro Lett.
2012, 1, 991.
[123] Y. P. Borguet, N. V. Tsarevsky, Polym. Chem. 2012, 3, 2487.
[124] A. Arbaoui, C. Redshaw, Polym. Chem. 2010, 1, 801.
[125] T. K. Goh, S. Yamashita, K. Satoh, A. Blencowe, M. Kamigaito,
G. G. Qiao, Macromol. Rapid Commun. 2011, 32, 456.
[126] G. Zhu, L. Zhang, Z. Zhang, J. Zhu, Y. Tu, Z. Cheng, X. Zhu,
Macromolecules 2011, 44, 3233.
[127] X.-B. Lu, D. J. Darensbourg, Chem. Soc. Rev. 2012, 41,
1462.
[128] J.-I. Kadokawa, S. Kobayashi, Curr. Opin. Chem. Biol. 2010, 14,
145.
[129] S. Kobayashi, in: Encyclopedia of Polymer Science and
Technology, John Wiley & Sons, New York 2010, p. 2/328.
[130] D. J. Xuereb, J. Dzierzak, R. Raja, Catal. Today 2012, 198, 19.
[131] M. K. Kiesewetter, E. J. Shin, J. L. Hedrick, R. M. Waymouth,
Macromolecules 2010, 43, 2093.
[132] R. A. Gross, M. Ganesh, W. Lu, Trends Biotechnol. 2010, 28,
435.
[133] J. E. Puskas, K. S. Seo, M. Y. Sen, Eur. Polym. J. 2011, 47, 524.
[134] M. T. Guzman-Gutierrez, D. R. Nieto, S. Fomine, S. L. Morales,
M. G. Zolotukhin, M. C. G. Hernandez, H. Kricheldorf, E. S.
Wilks, Macromolecules 2011, 44, 194.
[135] A. E. Hamielec, H. Tobita, in: Ullmanns Encyclopedia of
Industrial Chemistry, Wiley-VCH Verlag GmbH, Weinheim
2002, p. 1/133.
[136] Y. Jiang, J. M. J. Frechet, C. G. Willson, Polym. Bull. 1987, 17, 1.
[137] J. P. Montheard, M. Chatzopoulos, D. Chappard, J. Macromol.
Sci. R. M. C. 1992, 32, 1.
[138] S. Dumitriu, C. Dumitriu, in: Polymeric Biomaterials, S.
Dumitriu, (Ed., Marcel Dekker, New York 1994, p. 3/85.
[139] D. Y. Sogah, W. R. Hertler, O. W. Webster, G. M. Cohen,
Macromolecules 1987, 20, 1473.
[140] Y. Nagasaki, H. Ito, M. Kato, K. Kataoka, T. Tsuruta, Polym.
Bull. 1995, 35, 137.
[141] M. Kato, M. Kamagaito, M. Sawamoto, T. Higashimura,
Macromolecules 1995, 28, 1721.
[142] J. S. Wang, K. J. Matyjaszewski, J. Am. Chem. Soc. 1995, 117,
5614.
[143] T. E. Patten, K. J. Matyjaszewski, Adv. Mater. 1998, 10, 901.
n
~ ez, X. S. Wang, S. P.
[144] K. L. Robinson, M. A. Khan, M. V. D. Ba
Armes, Macromolecules 2001, 34, 3155.
[145] G. Masci, D. Bontempo, N. Tiso, M. Diociaiuti, L. Mannina, D.
Capitani, V. Crescenzi, Macromolecules 2004, 37, 4464.
[146] R. J. Zdrahala, I. J. Zdrahala, J. Biomater. Appl. 1999, 14, 67.

26

[147] P. A. Gunatillake, R. Adhikari, N. Gadegaard, Eur. Cells Mater.


2003, 5, 1.
[148] R. K. Kulkarni, K. C. Pani, C. Neuman, F. Leonard, AMA Arch.
Surg. 1966, 93, 839.
[149] B. M. Trost, Science 1991, 254, 1471.
[150] R. A. Sheldon, Green Chem. 2007, 9, 1261.
[151] R. A. Sheldon, Pure Appl. Chem. 2000, 72, 1233.
[152] P. W. Morgan, Condensation Polymers: by Interfacial and
Solution Methods, Wiley-Interscience, New York 1965.
[153] F. Millich, C. E. Carraher, J. J. Preston, Interfacial Synthesis,
Marcel Dekker, New York 1982.
[154] R. T. Mathers, S. P. Lewis, in: Green Polymerization Methods:
Renewable Starting Materials, Catalysis and Waste Reduction, R. T. Mathers, M. A. R. Meier, Eds., Wiley-VCH Verlag
GmbH, Weinheim 2011, p. 89.
[155] R. T. Mathers, K. C. McMahon, K. Damodaran, C. J. Reatarides,
D. J. Kelley, Macromolecules 2006, 39, 8982.
[156] J. E. McGrath, Ring-Opening Polymerization: Kinetics,
Mechanisms and Synthesis, ACS Symp. Ser. 286, American
Chemical Society, Washington D.C. 1985.
[157] R. H. Grubbs, Handbook of Metathesis, Wiley-VCH Verlag
GmbH, Weinheim 2003.
[158] R. Hoff, R. T. Mathers, Handbook of Transition Metal
Polymerization, John Wiley & Sons, Hoboken 2009.
[159] R. T. Mathers, K. Damodaran, M. G. Rendos, M. S. Lavrich,
Macromolecules 2009, 42, 1512.
[160] J. A. Laszlo, D. L. Compton, Biotechnol. Bioeng. 2001, 75, 181.
[161] Y. Liu, G. Wu, Radiat. Phys. Chem. 2005, 73, 159.
[162] F. Rantwijk, R. A. Sheldon, Chem. Rev. 2007, 107, 2757.
[163] L. X. Yang, Y. J. Zhu, W. W. Wang, H. Tong, M. L. Ruan, J. Phys.
Chem. B 2006, 110, 6609.
[164] M. Shadpour, D. Mohammad, Iran. Polym. J. 2010, 19, 983.
[165] H. Wang, Y. Liu, Z. Li, X. Zhang, S. Zhang, Y. Zhang, Eur.
Polym. J. 2009, 45, 1535.
[166] M. H. Fang, L. S. Wang, Int. J. Mol. Sci. 2007, 8, 470.
[167] S. Mallakpour, M. Dinari, e-Polymer 2007, 35, 1.
[168] J. Dupont, R. F. de Souza, P. A. Z. Suarez, Chem. Rev. 2002, 102,
3667.
[169] N. Jain, A. Kumar, S. Chauhan, S. M. S. Chauhan, Tetrahedron
2005, 61, 1015.
[170] J. Ding, D. W. Armstrong, Chirality 2005, 17, 281.
[171] J. Fraga-Dubreuil, M. H. Famelart, J. P. Bazureau, Org. Process
Res. Dev. 2002, 6, 374.
[172] K. Binnemans, Chem. Rev. 2005, 105, 4148.
[173] M. A. P. Martins, C. P. Frizzo, D. N. Moreira, N. Zanatta, H. G.
Bonacorso, Chem. Rev. 2008, 108, 2015.
[174] T. L. Greaves, C. J. Drummond, Chem. Rev. 2008, 108, 206.
[175] V. I. Parvulescu, C. Hardacre, Chem. Rev. 2007, 107, 2615.
[176] L. Liao, L. Liu, C. Zhang, S. Gong, Macromol. Rapid Commun.
2006, 27, 2060.
[177] E. I. Lozinskaya, A. S. Shaplov, Y. S. Vygodskii, Eur. Polym. J.
2004, 40, 2065.
[178] M. Yoneyama, Y. Matsui, High Perform. Polym. 2006, 18, 817.
[179] T. Ogoshi, T. Onodera, T. A. Yamagishi, Y. Nakamoto,
Macromolecules 2008, 41, 8533.
[180] P. Kubisa, Prog. Polym. Sci. 2009, 34, 1333.
[181] K. Bica, P. Gaertner, Eur. J. Org. Chem. 2008, 3235.
[182] J. F. Huang, H. Luo, C. Liang, D. Jiang, S. Dai, Ind. Eng. Chem.
Res. 2008, 47, 881.
[183] N. V. Tsarevsky, K. Matyjaszewski, Chem. Rev. 2007, 107,
2270.
[184] G. Imperato, B. Konig, C. Chiappe, Eur. J. Org. Chem. 2007,
1049.

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Applying the Principles of Green Chemistry to Polymer . . .


www.mre-journal.de

[185] J. L. Kendall, D. A. Canelas, J. L. Young, J. M. DeSimone, Chem.


Rev. 1999, 99, 543.
[186] P. Girotra, S. K. Singh, K. Nagpal, Pharm. Dev. Technol. 2013,
18, 22.
[187] J. A. Hyatt, Org. Chem. 1984, 49, 5097.
[188] M. A. McHugh, V. J. Krukonis, Supercritical Fluids Extraction:
Principles and Practice, 2nd edition Butterworth-Heineman,
Stoneham, MA, USA 1993.
[189] J. M. DeSimone, Z. Guan, C. S. Elsbernd, Science 1992, 257,
945.
[190] I. Yilgor, J. E. McGrath, V. Krukonis, Polym. Bull. 1984, 12, 499.
[191] V. Krukonis, Polym. News 1985, 11, 7.
[192] Z. Guan, J. R. Combes, Y. Z. Menceloglu, J. M. DeSimone,
Macromolecules 1993, 26, 2663.
[193] T. A. Hoefling, D. A. Newman, R. M. Enick, E. J. J. Beckman,
Supercrit. Fluids 1993, 6, 165.
[194] S. Salehpour, M. A. Dube, M. Murphy, Can. J. Chem. Eng.
2009, 87, 129.
[195] E. F. Jordan, Jr., B. Artymyshyn, J. Appl. Polym. Sci. 1969, 7,
2605.
[196] J. Hu, Z. Du, Z. Tang, E. Min, Ind. Eng. Chem. Res. 2004, 43,
7928.
[197] P. Cao, A. Y. Tremblay, M. A. Dube, Ind. Eng. Chem. Res. 2009,
48, 2533.
[198] N. T. McManus, G. Hsieh, A. Penlidis, Polymer 2004, 45,
5837.
[199] C. Quan, M. Soroush, M. C. Grady, J. E. Hansen, W. J.
Simonsick, Jr. Macromolecules 2005, 38, 7619.
[200] A. M. Stolin, A. G. Merzhanov, Polym. Eng. Sci. 1979, 19, 1065.
[201] M. Nomura, H. Tobita, K. Suzuki, D. I. Lee, Polymeric
Microspheres: Science and Technology, Kyoto University
Press, Kyoto, Japan 2007, p. 181.
[202] M. Goikoetxea, R. Heijungs, M. J. Barandiaran, J. M. Asua,
Macromol. React. Eng. 2008, 2, 90.
[203] S. Wang, Ph.D. thesis, Lehigh University, 2013.
[204] M. A. Dub
e, B. P. J. Soares, A. Penlidis, A. E. Hamielec, Ind. Eng.
Chem. Res. 1997, 36, 966.
[205] C. Decker, J. Coat. Technol. 1987, 59, 97.
[206] K. S. Anseth, S. M. Newman, C. N. Bowman, Biopolymers
1995, 22, 177.
[207] K. A. Berchtold, J. Nie, J. W. Stansbury, B. Hacioglu, E. R.
Beckel, C. N. Bowman, Macromolecules 2004, 37.
[208] L. G. Lovell, K. A. Berchtold, J. E. Elliott, H. Lu, C. N. Bowman,
Polym. Adv. Technol. 2001, 12, 335.
[209] H. Lu, J. W. Stansbury, C. N. Bowman, J. Dent. Res. 2005, 84,
822.
[210] J. W. Stansbury, C. N. Bowman, S. M. Newman, Phys. Today
2008, 61, 82.
[211] T. T. McMahon, K. C. Zadnik, Cornea 2000, 19, 730.
[212] H. M. Ma, R. H. Davis, C. N. Bowman, Macromolecules 2000,
33, 331.
[213] V. S. Khire, T. Y. Lee, C. N. Bowman, Macromolecules 2007, 40,
5669.
[214] C. G. Willson, B. C. Trinque, J. Photopolym. Sci. Technol. 2003,
16, 621.
[215] V. S. Khire, Y. Yi, N. A. Clark, C. N. Bowman, Adv. Mater. 2008,
20, 3308.
[216] K. S. Anseth, A. T. Metters, S. J. Bryant, P. J. Martens, J. H.
Elisseeff, C. N. Bowman, J. Controlled Release 2002, 78, 199.
[217] A. E. Rydholm, K. S. Anseth, C. N. Bowman, Acta Biomater.
2007, 3, 449.
[218] D. C. Neckers, Polym. Eng. Sci. 1992, 32, 1481.
[219] J. S. Young, S. R. Fox, K. S. Anseth, J. Manuf. Sci. E. -T. ASME
1999, 121, 474.

www.MaterialsViews.com

[220] R. R. Braga, R. Y. Ballester, J. L. Ferracane, Dent. Mater. 2005,


21, 962.
[221] C. N. Bowman, N. A. Peppas, Macromolecules 1991, 24, 1914.
[222] L. J. Gou, B. Opheim, C. N. Coretsopoulos, A. B. Scranton,
Chem. Eng. Commun. 2006, 193, 620.
[223] C. N. Bowman, C. J. Kloxin, AIChE J. 2008, 54, 2775.
[224] Y. Yagci, S. Jockusch, N. J. Turro, Macromolecules 2010, 43,
6245.
[225] J. Dou, Q. Zhang, M. Ma, J. Gu, J. Magn. Magn. Mater. 2012,
324, 3078.
[226] M. C. Rusu, C. Block, G. Van Assche, B. Van Mele, J. Therm.
Anal. Calorim. 2012, 110, 287.
[227] J. Jeong, B. Kim, S. Shin, B. Kim, J.-S. Lee, S.-H. Lee, J. K. Cho, J.
Appl. Polym. Sci. 2013, 127, 2483.
[228] Y. Liu, Y. Wang, J. Dent. 2013, 41, 71.
[229] J. Chiefari, Y. K. B. Chong, F. Ercole, J. Krstina, J. Jeffery, T. P. T.
Le, R. T. A. Mayadunne, G. F. Meijs, C. L. Moad, G. Moad, E.
Rizzardo, S. H. Thang, Macromolecules 1998, 31, 5559.
[230] R. T. A. Mayadunne, E. Rizzardo, J. Chiefari, J. Kristina, G. Moad,
A. Postma, S. H. Thang, Macromolecules 2000, 33, 243.
[231] C. Barner-Kowollik, M. Buback, B. Charleux, M. L. Coote, M.
Drache, T. Fukuda, A. Goto, B. Klumperman, A. B. Lowe, J. B.
Mcleary, G. Moad, M. J. Monteiro, R. D. Sanderson, M. P.
Tonge, P. Vana, J. Polym. Sci. Polym. Chem. 2006, 44, 5809.
ttgens, B. Klumperman, J. Meuldijk,
[232] R. Bussels, C. Bergman-Go
C. Koning, J. Polym. Sci. Polym. Chem. 2006, 44, 6419.
[233] B. Y. K. Chong, T. P. T. Le, G. Moad, E. Rizzardo, S. H. Thang,
Macromolecules 1999, 32, 2071.
[234] R. T. A. Mayadunne, J. Jeffery, G. Moad, E. Rizzardo,
Macromolecules 2003, 36, 1505.
[235] Y. Li, B. S. Lokitz, C. L. McCormick, Macromolecules 2006,
39, 81.
[236] B. Liu, A. Kazlauciunas, J. T. Guthrie, S. Perrier, Macromolecules 2005, 38, 2131.
[237] T. L. U. Nguyen, K. Eagles, T. P. Davis, C. Barner-Kowollik,
M. H. Stenzel, J. Polym. Sci., Part A: Polym. Chem. 2006, 44,
4372.
[238] G. Odian, Principles of Polymerization, 4th edition, John
Wiley & Sons, New York 2004.
[239] A. Sosnik, G. Gotelli, G. A. Abraham, Prog. Polym. Sci. 2011,
36, 1050.
[240] J. Li, X. Zhu, J. Zhu, Z. Cheng, Rad. Phys. Chem. 2007, 76, 23.
[241] Z. Cheng, X. Zhu, M. Chen, J. Chen, L. Zhang, Polymer 2003,
44, 2243.
[242] C. Costa, A. F. Santos, M. Fortuny, P. H. H. Arajo, C. Sayer,
Mater. Sci. Eng. C 2009, 29, 415.
[243] W. D. He, C. Y. Pan, T. Lu, J. Appl. Polym. Sci. 2001, 80, 2455.
[244] H. Zhao, H. Chen, Z. Li, W. Su, Q. Zhang, Eur. Polym. J. 2006, 42,
2192.
[245] D. Bogdal, P. Penczek, J. Pielichowski, A. Prociak, Adv. Polym.
Sci. 2003, 163, 193.
[246] R. Hoogenboom, U. S. Schubert, Macromol. Rapid Commun.
2007, 28, 368.
[247] M. P. Pollastri, W. G. Devine, in: Green Techniques for Organic
Synthesis and Medicinal Chemistry, W. Zhang, B. Cue, Eds.,
John Wiley and Sons, Chichester 2012, p. 325/342.
[248] A. Razavi, C. R. Acad. Sci. II C 2000, 3, 615.
[249] C. Bastiolo, Handbook of Biodegradable Polymers, iSmithers
Rapra Technology Limited, UK 2005.
chel, Biopolymers, Vol. 4, Polyesters III
[250] Y. Doi, A. Steinbu
Applications and Commercial Products, Wiley-VCH, Weinheim 2002.
[251] I. Kleeberg, C. Hetz, R. M. Kroppenstedt, R. M. Muller, W. D.
Decker, Appl. Environ. Microbiol. 1998, 64, 1731.

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

27

, S. Salehpour
M. A. Dube
www.mre-journal.de

nkel, S. Philipp, V. Reimer, K. O.


[252] M. Breulmann, A. Ku
Siegenthaler, G. Skupin, M. Yamamoto, in: Ullmanns
Encyclopedia of Industrial Chemistry, Wiley-VCH Verlag
GmbH, Weinheim 2009.
ller,
[253] I. Kleeberg, K. Wetzel, J. VandenHeuvel, R. J. Mu
Biomacromolecules 2005, 6.
[254] M. Mochizukim, in: Biopolymers. Polyesters III. Applications
chel, Eds., Wiley and Commercial Products, Y. Doi, A. Steinbu
VCH Verlag GmbH, Weinheim 2002, p. 1/23.
[255] A. Tullo, Chem. Eng. News 2002, 80, 13.
[256] K. S. Anderson, K. M. Schreck, M. A. Hillmyer, Polym. Rev.
2008, 48, 85.
[257] G. L. Baker, E. B. Vogel, M. R. Smith, Polym. Rev. 2008, 48, 64.
[258] K. D. Wendlandt, M. Jechorek, J. Helm, U. Stottmeiser, Polym.
Degrad. Stabil. 1998, 59, 191.
[259] S. Y. Lee, Biotechnol. Bioeng. 1996, 49, 1.
[260] J. Asrar, K. J. Gruys, in: Biopolymers. Polyesters III. Applications
chel, Eds., Wiley and Commercial Products, Y. Doi, A. Steinbu
VCH Verlag GmbH, Weinheim 2002, p. 53/81.
[261] S. F. Williams, D. P. Martin, in: Biopolymers. Polyesters III.
chel,
Applications and Commercial Products, Y. Doi, A. Steinbu
Eds., Wiley-VCH Verlag GmbH, Weinheim 2002, p. 91/27.
ller, U. Witt, E. Rantze, W. Deckwer, Polym. Degrad.
[262] R. J. Mu
Stabil. 1998, 59, 203.
[263] J. J. M. Swinkels, Industrial Starch Chemistry: Properties,
Modification and Applications of Starches, AVEBE, Veendam,
Netherlands 1999, p. 28.
[264] F. Kawai, in: Biopolymers. Miscellaneous Biopolymers and
chel,
Biodegradation of Polymers, S. Matsumura, A. Steinbu
Eds., Wiley - VCH Verlag GmbH, Weinheim 2002, p. 267/298.
[265] K. Ogata, F. Kawai, M. Fukaya, Y. Tani, Ferment. Technol.
1975, 53, 757.
[266] E. L. Fincher, W. J. Payne, Appl. Microbiol. 1962, 10, 542.
[267] F. Kawai, M. Fukaya, Y. Tani, K. Ogata, Ferment. Technol.
1977, 55, 429.
[268] Mowiol Polyvinyl Alcohol, Kuraray Product Brochure, www.
kuraray-poval.com/products/kuraray-poval-mowiol-mowiflex/,
accessed: June, 2013.
[269] A. Ashwin Kumar, K. Karthick, K. P. Arumugam, Int. J. Chem.
Eng. Appl. 2011, 2, 164.
[270] H. Tian, Z. Tang, X. Zhuang, X. Chen, X. Jing, Prog. Polym. Sci.
2012, 37, 237.
[271] G. E. Fonseca, M. A. Dube, A. Penlidis, Macromol. React. Eng.
2009, 3, 327.
[272] I. Alig, B. Steinhoff, D. Lellinger, Meas. Sci. Technol.
2010, 21.
[273] E. Frauendorfer, A. Wolf, W. D. Hergeth, Chem. Eng. Technol.
2010, 33, 1767.

28

[274] T. F. McKenna, S. Othman, G. Fevotte, A. M. Santos, H.


Hammouri, Polym. React. Eng. 2000, 8, 1.
[275] S. BenAmor, D. Colombie, T. McKenna, Ind. Eng. Chem. Res.
2002, 41, 4233.
[276] B. Alhamad, V. G. Gomes, J. A. Romagnoli, Int. J. Chem. React.
Eng. 2006, 4, 1.
mer, G. Niggemann, J. R. Leiza, J. M.
[277] R. Gesthuisen, S. Kra
Asua, Comput. Chem. Eng. 2005, 29, 349.
[278] J. Blomberg, P. J. Schoenmakers, U. A. T. Brinkman, J.
Chromatogr., A 2002, 972, 137.
[279] J. Beens, U. A. T. Brinkman, Analyst 2005, 130, 123.
[280] H. Tups, Chem-Ing-Tech 2010, 82, 531.
[281] S. Salehpour, M. A. Dube, Macromol. React. Eng. 2012, 6, 85.
[282] B. C. Smith, Fundamentals of Fourier Transform Infared
Spectroscopy, 2nd edition CRC Press, Taylor and Francis, New
York 2011.
[283] S. Shaikh, J. E. Puskas, Polym. News 2003, 28, 71.
[284] S. H. Patel, K. A. Bakeev, G. Chen, Q. Zhang, C. Wan, Adv.
Polym. Technol. 2010, 29, 1.
[285] S. Roberge, M. A. Dube, J. Appl. Polym. Sci. 2007, 103, 46.
[286] A. Al-Khanbashi, M. Dhamdhere, M. Hansen, Appl. Spectrosc.
Rev. 1998, 33, 115.
[287] M. Schmitt, J. Popp, J. Raman Spectrosc. 2006, 37, 20.
[288] M. Van Den Brink, J. F. Hansen, P. De Peinder, A. M. Van Herk,
A. L. German, J. Appl. Polym. Sci. 2001, 79, 426.
[289] Y. S. Kim, C. S. P. Sung, J. Appl. Polym. Sci. 1995, 57, 363.
[290] L. J. Bond, M. Morra, M. S. Greenwood, J. A. Bamberger, R. A.
Pappas, in Proceedings of the 20th IEEE Instrumentation and
Measurement Technology Conference, Vol. 2, Piscataway, NJ
2003, p. 1288/1293.
[291] S. Canegallo, M. Apostolo, G. Storti, M. Morbidelli, J. Appl.
Polym. Sci. 1995, 57, 1333.
ttmer, Meas. Sci. Technol.
[292] P. Hauptmann, N. Hoppe, A. Pu
2002, 13, R73.
[293] N. Gibson, Hazards X: Process Safety in Fine and Specialty
Chemical Plants, 1st edition, Symp. Ser. 115/The Institution of Chemical Engineers, Rugby, Warwickshire, UK
1989.
[294] J. Troitzsch, Plastics Flammability Handbook: Principles,
Regulations, Testing, and Approval, Hanser-Gardner Publications, Cincinnati, OH 2004, p. 3.
[295] P. Georlette, J. Simons, L. Costa, in: Fire Retardancy of
Polymeric Materials, A. F. Grand, C. A. Wilkie, Eds., Marcel
Dekker Inc., New York 2000, p. 245/284.
[296] V. Casson, G. Maschio, Macromol. Symp. 2011, 302,
273.
[297] A. L. Roes, M. K. Patel, Risk Anal. 2007, 27, 1311.

Macromol. React. Eng. 2014, 8, 728


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

You might also like