Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Chapter 1

Introduction to Polyurethane Chemistry


Felipe M. de Souza,1 Pawan K. Kahol,2 and Ram K. Gupta*,1

1Department of Chemistry, Kansas Polymer Research Center, Pittsburg State University,


Pittsburg, Kansas 66762, United States
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

2Department of Physics, Pittsburg State University, Pittsburg, Kansas 66762, United States

*Email: rgupta@pittstate.edu
Downloaded via 178.43.250.221 on November 6, 2022 at 16:34:10 (UTC).

Polyurethanes have improved the quality of modern life due to a large number
of applications such as insulation foams in houses and buildings to control
temperature, flexible foams in car seats and furniture, dampers in shoes for
comfort, and coatings to protect wooden and metallic materials or simply to
improve aesthetics, packing, and medical devices. Understanding the
structure–property relationship of polyurethane is a key element in designing new
materials with improved properties. Due to various requirements, ranging from
physical behavior to chemical properties to novel applications and sustainability,
polyurethane chemistries and technologies are rapidly responding to and adopting
such changes and needs. This chapter discusses some of the important topics of
polyurethanes including their chemistry, mechanism, and synthesis of new
materials (such as polyols and isocyanates). It also discusses the role and variety
of other ingredients—such as chain extenders, cross-linkers, catalysts, and
additives—to understand how they can influence the overall properties of
polyurethanes. Some examples of advances made in polyurethanes, and broader
perspectives about their chemistry and technology, are provided.

Introduction
Polyurethanes are an immensely versatile class of polymers used in insulators, foams, elastomers,
synthetic skins, coatings, adhesives, and so forth. Polyurethane was first developed through basic
diisocyanate polyaddition reactions by Dr. Otto Bayer and partners (1). In 1937, it reached
industrial-scale synthesis and it was established in the market in the 1950s. Polyurethanes are
characterized by the presence of a urethane linkage, and they can be easily synthesized through
an addition reaction between alcohol and an isocyanate. However, a functionality of at least 2 is
required in both the starting materials to prepare polyurethane. The reaction and the general formula
of polyurethane are given in Figure 1. Polyurethanes are widely explored because of their facile
synthesis that can be performed at room temperature and under mild conditions. The modifications

© 2021 American Chemical Society


Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
in the structure of polyols or isocyanate can be easily performed using various chemical approaches
that provide polyurethanes with a range of properties. For example, an elastic polyurethane can
be synthesized by using a polyol with a linear structure and high molecular weight with low
functionality. Meanwhile, a rigid polyurethane can be prepared by using a polyol with aromatic
groups in the structure, low molecular weight, and higher functionality for cross-linking (2). A
balance in chemical structure (hydrophobicity and hydrophilicity) provides water-dispersible
polyurethanes for coating applications.

Figure 1. General reaction for the synthesis of polyurethane through an addition reaction between a
bifunctional isocyanate and polyol.

Polyurethane is among the most important classes of industrial polymers with the fastest-
growing global market. This is due to the ability to modify the starting chemicals for a range of
polyurethanes with desired properties. The global polyurethane market was valued at more than
$65 billion and is expected to increase at a rate of 3.2% until 2027. When polyurethane foams are
used as a thermal insulator in housing, energy consumption decreases by 75% to 95%, creating
positive economic and sustainable effects (3). Their high electrical resistance allows application for
electrical insulation and dielectrics in electronic devices. The fact that polyurethanes are lightweight,
accompanied by their high mechanical strength, makes them a key component in the automotive
sector for improving efficiency due to less fuel consumption, increased safety, and enhanced comfort
of the vehicles. Flexible foams find a wide range of applications in the automobiles and furniture
industries where they are used, such as in beds, couches, and seat cushions. Flexible polyurethanes
are also used in the footwear industry as dampers in shoes and in the packing industry to protect
shipments from damage. Their chemical inertness and environmental stability make them very
suitable for use in adhesives and anticorrosion coatings. The latter has been receiving great attention,
as corrosion is one of the major issues in many countries and is responsible for an average gross
domestic product loss of 3% (4). Some of the novel applications of polyurethanes in areas like
biomedical and tissue engineering are possible as a result of their chemical stability, high surface area,
and flexibility. They can be used as synthetic scaffolds for the culture of cells or materials to aid in
the transport of fluids in the body, such as in synthetic veins or components of a pacemaker (5, 6).
The tunable properties of polyurethanes make them unique and suitable for a range of applications,
as mentioned previously.
Despite the large number of applications and huge market for polyurethanes, certain issues, such
as their high flammability, restrict their use in valuable applications and merit scientific attention.
Additionally, most of the chemicals used in the preparation of polyurethanes are derived from
petroleum-based resources, and there are growing concerns about sustainability and toxicity
associated with polyurethanes. The instability and degradation of polyurethanes in outdoor
applications also need attention. Researchers are using eco-friendly flame retardants in polyurethanes

2
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
to reduce their inherent flammability. They are applying newer approaches, such as non-isocyanate
routes, for polyurethanes to reduce their toxicity. They are also using renewable resources to
synthesize starting chemicals for polyurethanes for a sustainable future (7).

Materials and Chemistry of Polyurethanes


After arduous research, scientists understand how to optimize the reaction between hydroxyl
and isocyanate groups and how to control the properties of polyurethanes to make them suitable for
specific applications. This section covers the main materials used for the synthesis and formulation
of polyurethanes such as polyols, isocyanates, catalysts, blowing agents, surfactants, and additives.
From a general perspective, polyols are flexible segments of the polymeric chain that contribute
to the elasticity polyurethanes exhibit. When polyols are added in excess, the final polyurethanes
tend to be softer and more hydrophilic. Isocyanates are more reactive, are responsible for the curing
process, and lead to the rigid part of the polyurethanes—therefore, excess addition of isocyanate
provides more rigid and hydrophobic polyurethanes. Catalysts are frequently used in the preparation
of polyurethanes to enhance reactivity between polyols and isocyanates and to allow the reaction
to take place at low temperature (mostly at room temperature). Blowing agents, such as water, are
applied during the preparation of polyurethanes to provide foams. They also play an important role in
controlling the cellular structure and morphology by forming bubbles within polyurethanes during
the foaming process. Surfactants are responsible for improving the dispersion of the polyol and
isocyanate to ensure a homogenous system and to provide stable cellular structures by minimizing
the collapse of cells during the curing process. Sometimes, additives are incorporated in
polyurethanes to reduce the cost and provide specific applications such as flame retardancy, smoke
suppression during combustion, color, UV resistance, and mechanical strength.

Polyol Chemistry
By far, polyols compose the largest number of starting materials that can be synthesized to
design polyurethanes with a range of properties. They can be separated into two main groups: high
and low molecular weight polyols. Molecular weight, functionality, and nature of the polyols affect
the properties of the polyurethanes. About 70% of polyurethane produced worldwide accounts for
flexible/semiflexible foams, adhesives, coatings, elastomers, sealants, and other products, which are
mostly created from elastic polyurethanes. High molecular weight polyols, with long alkyl segments,
are used in preparing flexible or elastic polyurethanes as a result of their linear chains allowing
free rotation accompanied by low functionality (2–3) and low degree of cross-linking. Polyalkylene
oxide polyether polyols are the most frequently used polyols for flexible polyurethanes. They can be
synthesized through the self-polymerization of oxiranes (an addition reaction) that is initiated by a
polyol with low molecular weight (8). The reaction can be catalyzed by three groups of compounds:
anionic compounds, which include NaOH, KOH, Ba(OH)2, Sr(OH)2, C16H30CaO4, or
naphthenates; cationic compounds, which include BF3, CF3SO3H, PF5, or SbF5; and coordinative
compounds, which include a Al(OR)3, Zn(OR)2, Ti(OR)4, or Zn3[Co(CN)6]3.
Polyester polyols are the second most frequently produced polyols for polyurethanes. Polyester-
based polyurethanes are known to present properties such as higher crystallinity (due to stronger
intermolecular interactions between polymeric chains), higher thermal stability, and increased fire
resistance compared to polyether-based polyurethanes. They can also undergo hydrolysis when
exposed to moisture and heat for a long time. They can be synthesized through a polycondensation
reaction between polyols with dicarboxylic acids, or through a ring-opening reaction of cyclic

3
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
carbonates or lactones, as shown in Figure 2. The reaction requires a constant removal of water to
shift the equilibria toward the formation of polyester polyol. The reaction can be self-catalyzed due to
the acidity of carboxylic acid groups or through C7H8O3S, C16H30O4Sn, CH3COO-Zn+, or Mn2+,
which usually provide a better yield.

Figure 2. Chemical reactions for the synthesis of polyester polyols. Polycondensation through (a)
esterification, (b) transesterification, and (c) ring-opening.

Acrylic polyols are another group of polyols widely used for the preparation of polyurethanes.
Acrylic polyols are synthesized through a radical addition reaction at their unsaturation to form a
copolymer. At least one of the monomers must present a hydroxyalkyl group to react with isocyanate.
Figure 3a shows the chemical structure of some commonly used monomers for acrylic polyols.

4
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
The wide structural variety allows the design and implementation of different sets of properties.
For example, methylmethacrylate and styrene present a short structure that provides hardness and
water resistance, with the exception of methylmethacrylate being effective when exposed to light.
Another advantage is the presence of other reactive groups, such as carboxylic acids, which makes
methylmethacrylate readily soluble in water after neutralization. This feature creates proper
dispersion in polar solvents or water and low viscosity, which are important for ease of processing.
Some of the uses are for high-performance coatings in metals, as a result of proper adhesion or
formulation of waterborne polyurethanes (9, 10). Another aspect of the structure relates to the facile
implementation of larger dangling groups, such as butyl acrylate or 2-ethylhexyl acrylate, which adds
elasticity to the polymeric chain as a result of free movement and less packing. Figure 3b provides an
example of a radical addition copolymerization reaction. The reaction occurs in the presence of an
initiator and with the slow addition of the comonomer to control the structure of the final product.

Figure 3. (a) Common examples of acrylic monomer that can be used in the synthesis of acrylic polyols and
(b) an example of a radical addition copolymerization reaction for the synthesis of an acrylic polyol.

Polysiloxanes are another industrially important class of materials used in the preparation of
polyurethanes. Their most intriguing property is their low glass-transition temperature, which is
about −123 °C. In practice, polysiloxane-based polymers have a broad range of stable work
temperatures and are used as elastomeric materials. One of the important aspects of polysiloxane-
derived polyurethanes is that they maintain their elastic properties even at low temperatures. Also,
silicone-based polyurethanes are highly stable against aggressive environments, including electrical
currents, which allows their application in corrosion-resistive coatings and insulators.
Polyols with low molecular weight are another important group of starting chemicals used to
produce rigid polyurethanes. Their high functionality (~3–8) and short chains increase their viscosity

5
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
and lead to highly branched and cross-linked polyurethanes. Also, their reaction with aromatic
isocyanates, such as diphenyl methylene diisocyanate, provides rigidly structured polyurethanes. The
presence of many hydroxyl groups leads to a high number of urethane linkages, which increases
interactions among the chains. Rigid polyurethanes are generally obtained by using polyols with
short chains, high functionality, and aromatic segments in their structure, with more than 90%
closed cells. By contrast, flexible polyurethanes present an open-cell structure that allows free passage
of air. Rigid foams with around 90% closed cells do not allow air and heat passages and are used for
excellent thermal insulation in houses, buildings, and freezers. Both polyethers and polyester polyols
can be used for the preparation of rigid polyurethanes. However, because of chemical versatility,
lower cost, and higher viscosity, the use of the former is more common. The use of cross-linking
agents in polyurethanes is important to control some of the properties of polyurethanes, such as
rigidity and mechanical strength. The use of cyclic or aromatic cross-linkers improves packing and
decrease mobility, yielding structures that are mechanically strong, thermally stable, and resistant
to fire, compared with aliphatic cross-linkers. There are three types of cross-linkers for polyether
polyols, which are based on hydroxyl groups, amino groups, and condensates (i.e., phenols and
formaldehyde). Some examples of the starter polyols, along with their functionality, are given in
Figure 4.

Figure 4. Examples of some starter polyols along with their functionality.

6
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
Hydroxyl groups can be chemically attached to alkylene oxides through ring-opening, which
is the same as for amino-based cross-linkers. This can perform a nucleophilic attack at the partially
positive carbon, yielding an alcoholic group. The third type of cross-linker includes three main
compounds: Mannich bases, novolaks, and melamine-based polyols. Mannich bases are obtained
through an amino alkylation reaction. The base consists of a methylene bridge that originates from
formaldehyde and connects an alkyl amine with a phenol. Novolak resins are another case of
condensation reactions that occur without an alkylamine compound. Instead, the condensation
happens between phenol and formaldehyde. The third group is based on melamine polyols.
Melamine is a heterocyclic and aromatic substance that presents three amine groups. Despite its
high functionality, it is not an effective nucleophile for performing ring-opening into ethylene or
propylene oxide, due to its tautomeric forms. However, a known approach to functionalize it uses
formaldehyde and diethanolamine (or their derivatives) to obtain a Mannich base with high
functionality.
The extensive use of petrochemical-based materials for polyurethanes has created environmental
concerns and economic instability, triggering questions about how to use new sources that could
guarantee more stable and cheaper production. With more insight into these considerations, research
into renewable materials has yielded many promising biopolymers. Many bioderived starting
materials that naturally present hydroxyl groups—such as glycerol, sucrose, sorbitol, and ricinoleic
acid (derived from castor oil triglyceride)—can be used directly for polyurethane production.
Glycerol is a viable biosource for industrial applications because it is largely obtained as a byproduct
of corn oil, soybean oil, and biodiesel. Glycerol-based polyols are used in the production of flexible
polyurethanes, for application in mattresses, bedding, furniture, car seats, and so on. Other
biomaterials, such as vegetable and fruit oils, can be easily converted to polyols for polyurethanes.
The biopolyols can be synthesized through various methods, such as epoxidation followed by ring-
opening, ozonolysis, transesterification, and thiol-ene reactions. A few examples of starting materials
that have been converted to biopolyols are: lignin, terpenes, terpenoids, corn, and soybean oil.
With the help of available chemistries and materials as discussed previously, there are endless
possibilities to prepare polyurethanes with a large range of characteristics and applications. Materials
from both petrochemicals and renewable resources are heavily used for the industrial production
of polyurethanes. Many synthetic approaches are available to functionalize the starting materials,
such as introducing carboxylic groups to improve solubility in water; implementing flame-retardant
groups to reduce flammability; increasing the plasticizing effect through the extension of the main
chain; or using short-chain polyols to obtain a rigid structure. The large number of possibilities,
and the search for new technologies, are continuing to motivate the scientific community to further
explore polyurethanes for versatile applications.

Isocyanates and Their Characteristics


Isocyanates are characterized by the presence of reactive groups with the structure (-N=C=O).
This is the complementary reactive group for the synthesis of polyurethanes through an addition
reaction with hydroxyl. The number of commercial isocyanates available is much smaller compared
to polyols. There are a few major producers of isocyanates: Dow, Huntsman, Bayer, BASF, Shell,
ICI, and AC (1). Chemical structures of some of the commonly used isocyanates are shown in Figure
5. The importance of isocyanates is in their high reactivity toward hydroxyl groups, which makes
polyurethane production more efficient. One of the routes for the synthesis of isocyanates consists of
using amine-based compounds as starting materials, followed by a reaction with phosgene:

7
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
Two different types of isocyanates (aliphatic and aromatic) are generally used in the fabrication
of polyurethanes. Aromatic isocyanates provide rigid segments to the structures and impact the
overall properties of polyurethanes. For example, toluene diisocyanate (TDI) and methylene
diphenyl diisocyanate contain aromatic structures, and their applications mostly focus on rigid and
thermoset polyurethanes. With their relatively low cost and high reactivity compared to other
diisocyanates, these are the preferred isocyanates for industrial applications. One of the drawbacks
of aromatic isocyanates is their low stability in UV light. They tend to become brownish or yellowish
during prolonged exposure to UV light and therefore are largely used for indoor applications.
Aliphatic isocyanates, such as hexamethylene diisocyanate (HMDI) and isophorone diisocyanate
(IPDI). are more often used in coatings because they mix well with pigments, retain gloss aspect,
and are UV stabilized. Polyurethanes with isocyanate-terminated groups are generally hydrophobic,
chemically stable, and more rigid, while polyurethanes with hydroxyl-terminated groups are
hydrophilic, susceptible to external agents, and flexible (1, 2).

Figure 5. Common isocyanates for polyurethanes.

One of the main concerns about isocyanates is their toxicity. Under harsh conditions, such as
high temperatures, they usually release voltaic organic components that are known to cause health
problems. One of the solutions to this problem is to use polymeric isocyanates, which have high
vapor pressures. Other approaches are focused on finding new synthetic routes that do not use
isocyanates for polyurethanes. As an example, limonene was used to prepare polyurethane with a
non-isocyanate route in a three-step process (11). In the first step, epoxidation of the double bond
of limonene was performed. In the second step, limonene dioxide was put into a reaction with CO2
under high pressure, and in the presence of a catalyst, to insert a carbonate group in the structure of
limonene dioxide. In the final step, the amine-based compound was used to react with the carbonate
to create urethane linkage. The schematics of this process are given in Figure 6.

8
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
Figure 6. Isocyanate-free route for the synthesis of polyurethane through epoxidation. This is followed by the
introduction of CO3-2 with the reaction of CO2, which is then followed by a reaction with an amine.
Reproduced with permission from reference (11). Copyright 2012 Royal Society of Chemistry.

Reactions of Polyols and Isocyanates


Wide commercial applications of polyurethanes are possible as a result of the high reactivity of
isocyanates with polyols. The resonance structure of the isocyanate is displayed in Figure 7. Due
to electronegativity difference, while negative partial charges are generated at both nitrogen and
oxygen, the carbon center becomes positively charged. It is susceptible to nucleophile attacks with
active hydrogen (i.e., R-OH). The reactivity of the isocyanates can be increased if R is an electron-
withdrawing group such as an aromatic group. This explains the high reactivity of commonly used
isocyanates in the industry, such as methylene diphenyl diisocyanate and TDI, compared to aliphatic
isocyanates such as HMDI and IPDI.

Figure 7. Resonance structure of isocyanate.

The reactivity of hydroxyl groups in polyols is an important factor in designing new experiments
for specific applications, such as spray foaming, where high reactivity and fast curing time are desired.
The order of reactivity of the hydroxyl groups in the polyol is: primary hydroxyl groups are more
reactive than secondary hydroxyl groups, which are more reactive than tertiary hydroxyl groups,
which are more reactive than aromatic (phenol) hydroxyl groups. In polyurethanes, while the rigid
domains are composed of urethane linkages that can interact with each other through hydrogen
bonding, the soft domains are derived from a polyol, which contains longer chains. Although the
reaction between the hydroxyl and isocyanate groups is fast, phase separation often occurs due to
differences in polarity and density of the isocyanates and polyols. Surfactants are therefore often

9
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
used in polyurethanes to promote a homogenous phase. Catalysts are also used in the synthesis
of polyurethanes to enhance reaction rates. Organotin and tertiary amine-based compounds are
commonly used as catalysts for polyurethanes. Although the mechanism of the activation reaction
between isocyanate and alcohol is not fully established, Bloodworth and Davies proposed a route that
is the most accepted mechanism to date (12). The proposed mechanism suggests that the nitrogen
from isocyanate coordinates with the organotin, forming the intermediate N-stannylurethane, which
then induces the alcoholysis. The detailed mechanism is shown in Figure 8 (13).

Figure 8. Mechanism for urethane linkage through an addition reaction of isocyanate, with a hydroxyl
group catalyzed by organotin. Reproduced with permission from reference (13). Copyright 2013 American
Chemical Society.

An activation reaction can also occur in the absence of a catalyst where the nucleophilic oxygen
from the alcoholic group can react with the partially positive carbon in the isocyanate. It then leads to
a proton transfer from the oxygen of the alcohol to the nitrogen of the isocyanate, yielding a urethane
bond. This reaction is described in Figure 9a. This reaction can also be catalyzed by tertiary amines.
The key element for the catalysis is the presence of a free pair of electrons in the nitrogen from the
tertiary amines, which polarize the hydrogen atom from the alcohol, prompting the oxygen from the
alcohol to perform a nucleophilic attack on the partially positive carbon from the isocyanate group.
An intermediate structure is created, which leads to the urethane linkage. This mechanism is shown
in Figure 9b. Strong Lewis or Brønsted acids, such as organotin, can be also used as catalysts for this
reaction (14). Other types of catalysts are tertiary amines that increase the reaction rate of isocyanates
with water (used as a reactive blowing agent) to make polyurethane foams (15). In this mechanism,
acid acts on the free electron pair of the nitrogen from the isocyanate. This polarizes the bond and
allows the electrophilic carbon from the isocyanate to be attacked by the oxygen from the alcohol.
This is followed by a proton transfer step that leads to the urethane linkage, as shown in Figure 9c.

10
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
Figure 9. Mechanism routes of a urethane addition reaction between isocyanate and hydroxyl groups (a)
catalyst-free, (b) through base catalysis (nucleophilic activation), and (c) through acid catalysis
(electrophilic activation). Reproduced with permission from reference (14). Copyright 2015 American
Chemical Society.

Various proposed mechanistic routes for the synthesis of polyurethanes provide innovative ideas
to scientists in the design of new experiments that are cost-effective and facile. One of the approaches
is a one-step process where polyurethanes can be rapidly prepared. In this approach, a homogenous
mixture containing all the reagents—such as polyols, surfactants, blowing agents, catalysts, cross-
linkers, and chain extenders except the isocyanate—is prepared. This is followed by the rapid
addition of the required amount of isocyanate. The quality of polyurethane foams depends on some
experimental factors, such as the amount of surfactants and catalysts, mixing speed, and stirring time
(8). The reaction between isocyanate and hydroxyl groups can be performed through facile methods,
under mild conditions, and with low-cost instruments.

Other Materials Used for Polyurethanes


Polyols and isocyanates are the core components in the production of polyurethanes. The
chemical structure and functionality of polyols and isocyanates are expected to affect the properties
of polyurethanes. However, other materials used in polyurethanes can also influence properties such
as density, porosity, morphology, mechanical strength, and rate of reaction. Material properties
are also affected by other factors such as catalysts, chain extenders, cross-linkers, surfactants, and
blowing agents.
Catalysts are substances that have the ability to decrease the activation energy of a reaction
and therefore increase its rate. They carry great economic importance by improving the speed of
production and enabling faster manufacturing and curing processes. As previously mentioned, metal
complexes are used as catalysts for the urethane formation reaction. This can occur through various
organometallic compounds based, for example, on antimony, tin, bismuth, zinc, lead, and mercury.

11
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
With the toxicity of many of these compounds, bismuth and zinc are receiving more attention as
safer catalysts. Other groups of catalysts are tertiary amine-based catalysts such as 1,4-
diazabicyclo[2.2.2]octane, triethylenediamine, dimethylethanolamine, and
dimethylcyclohexylamine. Chemical structures of some common catalysts are given in Figure 10.
Chain extenders and cross-linkers used in polyurethanes are additives of low molecular weight
that affect morphology, cellular structure, and thermal and mechanical properties. As a result of
their short chains with low mobility, chain extenders are amine- or hydroxyl-terminated compounds
with a functionality of 2 and with the role of introducing rigid domains in the polymeric structure.
Accompanied by the rigidity of the isocyanate groups, the use of chain extenders can decrease the
elasticity of the polyurethanes, which can be adjusted as desired. Cross-linkers with a higher
functionality (3 or more) create a covalent bond between the polymeric chains and lead to a densely
interconnected network. Cross-linked polyurethanes are mechanically stronger than the equivalent
linear chain–based polyurethanes because of the presence of chemical bonds instead of weaker
intermolecular interactions. Cross-linking also reflects on the thermal properties. Linear
(thermoplastic) polyurethanes (TPUs) become fluid at a temperature higher than their glass-
transition temperature. Meanwhile, highly cross-linked (thermoset) polyurethanes decompose at
higher temperatures instead of becoming fluid. The higher amount of energy required to disrupt
chemical bonds in thermoset polyurethanes makes them thermally stable.

Figure 10. Most employed catalysts for the synthesis of polyurethanes.

Surfactants are components that are added during the formation of polyurethanes to guarantee
proper mixing of the polyol and isocyanate. This is because of their inherent poor interaction with
each other. Surfactants provide the stability in growing cellular structures. Most of the surfactants
are either silicon-based block polymers, such as polydimethylsiloxane–polyoxyalkylene, or silicon

12
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
oils like nonylphenol ethoxylates. During the foaming process, they are used to form an emulsion
and control the cell structure. This is important in preventing imperfections, such as holes both
in and outside the foam. They are also used for nonfoaming purposes, allowing the release of air
and acting as a wetting agent, which is important for the development of coatings to evenly cover a
surface and avert delamination. Surfactants are classified into two categories based on their character;
they are either cationic or nonionic. The first group of surfactants is known to be more effective for
anticorrosion coatings, while the second group is more effective in decreasing superficial tension and
allowing proper mixing of the components.
Blowing agents are the compounds that are added during the formulation of polyurethanes to
provide foam structure by decreasing density and increasing porosity. They can be also divided into
two groups: physical and chemical blowing agents. Physical blowing agents are gases directly blown
into the polyurethane mixture during the foaming process. The chlorofluorocarbons (CFCs), such
as trichlorofluoromethane and dichlorodifluoromethane, are examples of physical blowing agents
that were largely used during the 1950s. Their application as blowing agents provides foams with
low-density, high-thermal, and electrical-insulating behavior; high closed-cell content; and good
mechanical properties. This makes them superior to other blowing agents. However, it was later
discovered that CFCs ascend to the stratosphere and undergo photodecomposition to produce
chlorine radicals that catalyze the decomposition of ozone into oxygen. This leads to the enhanced
greenhouse effect. The international regulation organization has since created the Montreal and
Kyoto Protocols, which demand the replacement or drastic decrease of CFCs’ usage. As an answer
to the strict requirements, the hydrochlorofluorocarbons were developed as a transition state for
the CFCs. Some examples of these compounds are CH2FCF3, CF3CH2CHF2, CF3CH2CF2CH3,
and CHF2CH3. Due to the lesser content of chlorine that could form radicals when exposed to UV
light, hydrochlorofluorocarbons were less aggressive compared with CFCs. Other blowing agents,
known as volatile organic compounds (VOCs), were later implemented as alternative options, but
they ascend in the atmosphere and contribute to the enhanced greenhouse effect. The difference
with them is that photodecomposition occurs in the lower atmosphere. This process causes a
phenomenon known as photochemical smog, which is usually observed in big cities with higher
concentrations of VOCs; high flammability is another concern. Strict regulations are therefore
required for the emission of these compounds. Some examples of VOCs are cyclopentane, n-
pentane, isopentane, and isobutane. Other options include methyl chloroform, acetone, and carbon
dioxide. The other major group is chemical blowing agents. These compounds undergo a chemical
decomposition leading to the formation of gases, which causes the blowing effect. Water is the
simplest and most frequently used chemical blowing agent, with its high reactivity with isocyanate
leading to the formation of CO2 (16).

Applications of Polyurethanes
With versatility in the chemistry and constituent materials, polyurethanes find ubiquitous
applications in virtually all fields. In the following sections, some of the important applications of
polyurethanes are briefly discussed to give an overview to readers.

Polyurethanes in Households and Construction


Building and construction sectors greatly rely on the use of rigid and flexible polyurethanes for
thermal insulation and applications in furniture, due to their suitable properties such as low density,
high thermal insulation, and mechanical ability. Highly reactive polyurethanes can be prepared by

13
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
selecting proper constituents and can be used for spray foams to cover most of the surfaces and
shapes. They are among the best thermal insulators to decrease energy consumption from running
air conditioning and heating. This helps the environment and provides more comfort. One work
reported a novel approach for the preparation of polyurethanes with efficient thermal insulating
behavior (17). These high-performance polyurethanes consisted of TPUs sandwiched between two
layers of polyimide (PI) film (Figure 11). A unique architectural structure was obtained by using
CO2, which allowed the polymer to have good elasticity and a thermal barrier. The thermal insulation
properties of the TPU film dropped from 0.257 to 0.162 W·(m·K)-1. This is a 37% change in the
insulating characteristics with this unique architectural design.

Figure 11. Scheme for the preparation of insulator TPU film: (a) premold system for the TPU pellets in
between the PI films, (b) compression of the sandwiched system with a pressure of 15 MPa under 190 °C
around 3 min, (c) saturation of CO2 that could flow within the TPU as represented by the arrows, and (d)
and (e) removing the PI film to obtain the thermal insulator TPU film. Reproduced with permission from
reference (17). Copyright 2018 American Chemical Society.

Polyurethanes in the Automotive Industry


The automotive industry uses flexible polyurethanes for car seats and rigid polyurethanes for
thermal and sound insulations. Without question, the most important features for polyurethanes
in vehicles are low weight accompanied by high mechanical strength. These features improve the
mileage, cost-efficiency of the fuel, and safety against collisions (18, 19). Polyurethanes are also used
in automobile coatings. Coatings are important for automobiles, as they provide corrosion resistance
to metals used in the body parts. They also provide a gloss effect to make automobiles weather
resistant, durable, and attractive. Automobile and furniture industries use flame retardants in their
coatings for added safety. One work studied the presence of flame retardants and their effect on car
dust (20). 2,2-bis(chloromethyl)-propane-1,3-diyltetrakis(2-chloroethyl) bisphosphate, known as
V6, is used as a flame retardant in automobile foam, which contains tris(2-chloroethyl) phosphate as
a known carcinogen compound (Figure 12). A concentration in the range of 5–6160 ng/g of V6 in
car dust was observed, which was much higher than in house dust. Although halogen-based flame

14
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
retardants are effective in quenching fire, their toxicity from the release of carcinogenic gases is a
major drawback. A fair amount of research has been devoted to synthesizing new materials that are
efficient flame retardants without the level of toxicity shown by halogen-based flame retardants. The
majority of the materials that have been used as green flame retardants are based on metal oxides (21),
nitrogen (22), phosphorus (23), and carbon (24). Aluminum trihydroxide, melamine, melamine
cyanurate, melamine phosphate, ammonium phosphate, red phosphorus, phosphate esters,
phosphinates, phosphonates, carbon black, and expandable graphite are a few examples of viable
and eco-friendly retardants. It is abundantly clear that the development and study of flame
retardants—which present proper compatibility with polyurethanes and do not produce toxic smoke
during the combustion process—are significantly important.

Figure 12. Chemical structure of two toxic flame retardants mostly used in car seats and furniture.
Reproduced with permission from reference (20). Copyright 2013 American Chemical Society.

Polyurethanes in Marine Life


Similar to the corrosion prevention applications in automobiles, the use of polyurethanes in
marine applications for boats, ships, oil platforms, and other boat-related technologies is very
important. Water in the ocean and seas is highly corrosive from the presence of large concentrations
of dissolved salts. The development of anticorrosion coatings that can endure such an aggressive
environment is desired. The waterborne polyurethanes have conquered a large share of the market in
the area of corrosion-resistive coatings, as a result of their eco-friendly nature, high adhesion to the
surface, and environmental stability. Their structure is made up of a polarized segment incorporated
into the main backbone, which improves their dispersion in water. This allows coatings to properly
disperse in polar solvents. This structure addresses an important environmental issue, as it drastically
reduces the use of VOCs. The implementation of other materials to improve corrosion resistance
has become a hot research area (25–28). Many composite coatings of waterborne
polyurethanes—functionalized with silicon, graphene, metal oxides, conductive polymers, and
nanomaterials—have been developed. The list of requirements for these composites contains proper
adherence to the substrate; mechanical resistance against scratching; abrasion and impact; and
chemical resistance against solvents, acids, bases, and radiation. As an example, a group of
researchers has synthesized reduced graphene and covalently grafted it with IPDI to obtain a

15
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
functionalized graphene–IPDI isocyanate (IP-GO). Then, it was dispersed with sodium
dodecylbenzene sulfonate, and a waterborne polyurethane (WPU) was added to percolate the IP-
GO into its polymeric chains to prepare IP-GO/WPU (29). The schematic for this process is shown
in Figure 13.

Figure 13. Schematic of the synthetic procedure for graphene oxide being functionalized with IPDI to obtain
IP-GO through (a) graphite, (b) graphene oxide, (c) IP-GO, and (d) a composite coating adhered to a
metal substrate. Reproduced with permission from reference (29). Copyright 2019 American Chemical
Society.

Wen and coworkers applied a composite coating over a metal surface and investigated its
efficiency as an anticorrosion coating by making measurements using the salt spray test (29). The
effects of salt spray can be observed in Figure 14. It was notable that IP-GO/WPU presented the
most efficiency during the salt spray test. One of the primary reasons for this is believed to be the
broad area that the graphene covered, which improved the physical shielding effect and prevented the
permeation of corrosive species. Also, proper dispersion of graphene into the WPU matrix played
an important role in preventing delamination and made a stable suspension for the coating. Effective
percolation of IP-GO into the WPU aided in preventing the corrosion.

16
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
Figure 14. Images for the samples coated with neat WPU, graphene oxide dispersed in the WPU, reduced
graphene oxide dispersed in the WPU, and IP-GO/polyurethane composite coating. The test was performed
for 480 h. Reproduced with permission from reference (29). Copyright 2019 American Chemical Society.

Polyurethanes for Coatings


“Coatings” is a very general term that describes many types of materials with many applications,
as they can be developed for anticorrosion, scratch protection, UV resistance, gloss effect, aesthetics,
pigments, hydrophobicity, flame retardancy, antibacterial properties, and many other purposes. A
group of scientists developed a partially green coating by synthesizing a hyperbranched polyurethane
using an A2 + B3 scheme, where TDI was the A2, castor oil–based polyol was the B3, 1,4-butanediol
was the chain extender, and poly(ε-caprolactone)-diol (PCL) was the macroglycol (30). The final
coating presented effective properties, such as high elongation at break reaching around 791%, a
tensile strength of 7 MPa, a scratch hardness of 4.5 kg, and relatively high chemical resistance and
hydrophobicity. This showed that such biobased materials can create effective properties for coatings.
Other reports have demonstrated an interesting synthetic route to obtain a self-healing WPU that
functions on the concepts of degradation and reassembly of disulfide bonds (31). The WPU was
obtained by using polypropylene glycol as the polyol, IPDI as the isocyanate, and 2-
bis(hydroxymethyl) propionic acid as the polar segment that introduces solubility to WPU. Then,
1,4-butylene glycol was used as the chain extender. The disulfide source, which grants the self-
healing properties, was 2-hydroxyethyl disulfide. The results demonstrated that an optimum amount
of disulfide bonds was able to fully fix the crack when exposed to 75 °C for 15 min.
Another route to obtain self-healing WPU was developed by taking advantage of the reversibility
of a Diels–Alder reaction system (32). The material was able to fix a crack when exposed to the
thermal curing process either from 65 °C for 24 h or 130 °C for 30 min, with an efficacy of around
93% (Figure 15). The self-healing properties were from the presence of the Diels–Alder diol, which
also increased the thermal stability of the polyurethane. This provided an increase in its degradation

17
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
to 50% from 333 to 357 °C. Another important property the WPU composite acquired was its
recyclability, which allowed this material to be recasted and reshaped in a hot pressing process.

Figure 15. The self-healing mechanism and recycling process for the WPU. Reproduced with permission
from reference (32). Copyright 2018 American Chemical Society.

Biomedical Applications of Polyurethanes


Polyurethanes are widely used in biomedical applications such as artificial skin, hospital
bedding, dialysis tubes, pacemaker components, catheters, and surgical coatings. The
biocompatibility, mechanical properties, and low cost are major factors to the success of
polyurethanes in the medical field.
The development of implants usually requires a high content of biobased components, because
the body rejects them less. In the case of polyurethanes, the biocomponent can vary from 30 to
70%, which creates a broader scope for applications in such areas (2). The biobased polyurethanes
are increasing their market share and are expected to reach about $42 million by 2022, which is
a minuscule percentage of the overall polyurethane market (less than 0.1%). Nonetheless, it is a
promising area, and intensive research is ongoing concerning the use of more biobased materials in
polyurethanes. Improvement is needed in the properties of biobased polyurethanes to match the
existing requirements, in order to scale up investment.
Biobased crystalline polyurethane was synthesized via a reaction of PCL, HMDI, and water that
played the role of a chain extender (33). Degradation tests were performed to study the stability of
biopolyurethane in simulated body fluids, such as phosphate-buffered saline solution. The changes
in thermal, mechanical, and physical properties were analyzed and compared to the equivalent
polyurethane obtained through using ethylene glycol as a chain extender instead of water. The results
demonstrated that the polyurethane obtained using water as a chain extender presented better
properties over time compared with its petrochemical equivalent. This not only greatly decreases
the cost of the process, but it also provides a facile route to obtaining value-added medical materials
that are suitable for joint endoprostheses (33). This was followed by another approach based on
this concept, which synthesized a biopolyurethane urea by utilizing rapeseed oil–based polyol, PCL,

18
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
HMDI, and water as a chain extender (6). To increase the surface area, sodium chlorine was used
to improve the porosity of the prepared polymers. The synthesized polymer was used as a scaffold
due to its porous structure to induce cell growth of the bone tissue. With similar results compared
to the previous example, the polyurethane that was exposed to simulated body fluid presented high
stability, providing a viable option for scaffold applications. Polyurethane ionomers are another
interesting class of polymers used for biomedical applications, as a result of their biocompatibility
and proper interaction with the body environment. Polyurethane ionomers can be used as tube
components for pacemakers and hemodialysis (34, 35).
The development of an effective drug delivery system is an important research area that is
currently focused on finding ways to tackle down cancer. An amphiphilic nanoparticle of
polyurethane based on L-lysine was prepared for drug delivery applications (36). This nanocarrier
was effectively loaded with doxorubicin, which is an effective drug treatment for cancer cells (Figure
16). The hydrophobic segments of the polyurethane interacted with the drug, and the hydrophilic
segments interacted with the cells. This system created a core-shell structure through a self-assembly
mechanism and was able to efficiently deliver drugs via two routes. First, the thermal response of
the nanoparticle acted as a trigger in releasing the drug at the cancer cell’s temperature (~41–43
°C), which is an extracellular response. Second, the aliphatic segments of the polyurethane suffered
enzymatic biodegradation by the action of lysosomes, allowing doxorubicin to be released inside the
cancer cell; this is an intracellular response. More than 90% of the breast cancer cells were killed,
while low cytotoxicity was maintained for healthy cells.

Figure 16. Overall scheme for the drug delivery system based on an amphiphilic polyurethane nanoparticle
to target cancer cells. Reproduced with permission from reference (36). Copyright 2019 American Chemical
Society.

Other Applications of Polyurethanes


Polyurethanes are heavily used in the packing sector due to their lightweight, low-cost, and
high shock absorption properties. The clothing industry uses polyurethanes to give elasticity to
spandex fibers and to improve thermal insulation in leathers. They are also used as finishing agents
to provide softness and easy cleaning of clothes (37). Composite coatings based on waterborne
polyurethanes—and functionalized with carbon-based materials such as activated carbon, graphene,

19
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
and fullerene—are used to shield electromagnetic interference to protect sensitive electronic devices
(38). The π–π interaction originating from the aromatic structure of these materials is known to
shield electromagnetic interference. Therefore, a broad range of applications is possible using
polyurethanes. Self-healing polyurethanes have been prepared using a 3-D printer (39). The
development of polyurethane acrylate with disulfide bonds in the structure is the primary reason for
its self-healing properties. Aside from the self-healing properties, other characteristics were notable,
such as: more than 400% elongation at break with a tensile strength of 3.4 MPa and self-healing
efficiency of 95% under 80 °C for half a day. Digital photographs of the prepared polyurethanes
with self-healing properties are shown in Figure 17. The novel use of 3-D printers in preparing
polyurethanes opens new possibilities for advanced applications in sensors, elastomers, flexible
electronics, and many other devices of various sizes and shapes—even in complicated structures
where these applications may not be possible with conventional synthetic and processing methods.

Figure 17. 3-D printed self-healing polyurethane acrylate containing disulfide bonds. Reproduced with
permission from reference (39). Copyright 2019 American Chemical Society.

As previously discussed, porous polyurethanes can be used as scaffolds for tissue engineering.
Using a different approach, a group of researchers used polyurethanes as an adhesive to join broken
bones by inducing their growth with a low amount of rejection by the body (40). The adhesive
strength was varied by increasing the isocyanate and β-tricalcium phosphate, which was used as a
ceramic filler. The adhesion was twice as strong as in the commercially available bone cement, which
was based on poly(methyl methacrylate). The porosity could be controlled based on the content
of added water during the formulation. The synthetic process for a bone adhesive polyurethane
is shown in Figure 18. Biobased polyurethanes for connecting living tissue were also developed,
as shown in Figure 19 (41). The synthesis consisted of using sugar xylose as a cross-linker and
biodegradable segment, along with polyethylene glycol and cyclohexyl isocyanate. The results
demonstrated effective adhesion properties, reaching 95 kPa for muscle tissue and 415 kPa for an
aluminum substrate with 15% of xylose added to the formulation. Biodegradation of this biobased

20
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
polyurethane was found to be 20% after 2 months. The high adhesiveness, biodegradability, and
nontoxicity of such polyurethanes diversify their applications in biomedical research.

Figure 18. Synthesis for a bone adhesive polyurethane. Reproduced with permission from reference (40).
Copyright 2019 American Chemical Society.

Figure 19. Schematics for a polyurethane-based biomedical adhesive connecting living tissue. Reproduced
with permission from reference (41). Copyright 2016 American Chemical Society.

Concluding Remarks
Polyurethanes are among the most versatile classes of polymers, and they are generally
synthesized using addition reactions of polyols and isocyanates. A wide range of materials and applied
facile chemistries are capable of providing polyurethanes with a variety of properties and
applications. Some characteristics—such as high thermal and electrical insulation, low weight,
rigidity, and flexibility—allow these polymers to be used as core components in insulation, paints,
coatings, elastomers, foams, and so forth. Over the last decade, extensive research has been done to
understand their sustainable chemistry, especially in biopolyurethanes, but more work is needed to
make polyurethanes green and sustainable.

References
1. Janik H., Sienkiewicz M., Kucinska-Lipka J. Polyurethanes. In Handbook of Thermoset Plastics,
Third ed.; Dodiuk H., Goodman, S. H., Eds.; William Andrew Publishing: Boston, 2014; pp
253–295.

21
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
2. Akindoyo, J. O.; Beg, M. D. H.; Ghazali, S.; Islam, M. R.; Jeyaratnam, N.; Yuvaraj, A. R.
Polyurethane types, synthesis and applications-a review. RSC Adv 2016, 6, 114453–114482.
3. Polyurethanes. https://www.polyurethanes.org/en/sustainability/energy-efficiency/
buildings-and-the-passive-house (accessed Dec 1, 2020).
4. Koch G., Varney J., Thopson N., Moghissi O., Gould M., Payer J. International Measures of
Prevention, Application, and Economics of Corrosion Technologies Study, 2016. http://impact.
nace.org/documents/Nace-International-Report.pdf (accessed Nov 17, 2020).
5. Remya, V.; Patil, D.; Abitha, V.; Ajay Rane, R. M. Biobased materials for polyurethane
dispersions. Chem Int 2016, 2, 158–167.
6. Zieleniewska, M.; Auguścik, M.; Prociak, A.; Rojek, P.; Ryszkowska, J. Polyurethane-urea
substrates from rapeseed oil-based polyol for bone tissue cultures intended for application in
tissue engineering. Polym Degrad Stab 2014, 108, 241–249.
7. Babb D. A. Polyurethanes from Renewable Resources. In Synthetic Biodegradable Polymers;
Babb D. A., Ed.; Springer: Verlag Berlin, 2011; pp 315–360.
8. Ionescu, M. Mihail Ionescu: Chemistry and Technology of Polyols for Polyurethanes. Polimeri
2006, 26, 218–218.
9. Brinkman, E.; Vandevoorde, P. Waterborne two-pack isocyanate-free systems for industrial
coatings. Prog Org Coatings 1998, 34, 21–25.
10. Xu, H.; Yang, D.; Guo, Q.; Wang, Y.; Wu, W.; Qiu, F. Waterborne Polyurethane-Acrylate
Containing Different Polyether Polyols: Preparation and Properties. Polym Plast Technol Eng
2012, 51, 50–57.
11. Bähr, M.; Bitto, A.; Mülhaupt, R. Cyclic limonene dicarbonate as a new monomer for non-
isocyanate oligo- and polyurethanes (NIPU) based upon terpenes. Green Chem 2012, 14,
1447–1454.
12. Bloodworth, A. J.; Davies, A. G. Organometallic reactions. Part I. The addition of tin alkoxides
to isocyanates. J Chem Soc 1965, 9750, 5238–5244.
13. Delebecq, E.; Pascault, J-P; Boutevin, B.; Ganachaud, F. On the Versatility of Urethane/Urea
Bonds: Reversibility, Blocked Isocyanate, and Non-isocyanate Polyurethane. Chem Rev 2013,
113, 80–118.
14. Sardon, H.; Pascual, A.; Mecerreyes, D.; Taton, D.; Cramail, H.; Hedrick, J. L. Synthesis of
Polyurethanes Using Organocatalysis: A Perspective. Macromolecules 2015, 48, 3153–3165.
15. Sonnenschein M. F. Polyurethanes: Science, Technology, Markets, and Trends; John Wiley &
Sons: Hoboken, NJ, 2014.
16. Singh S. N. Blowing Agents for Polyurethane Foams; Rapra Technology: Hamburg, Germany,
2001.
17. Ge, C.; Zhai, W. Cellular Thermoplastic Polyurethane Thin Film: Preparation, Elasticity, and
Thermal Insulation Performance. Ind Eng Chem Res 2018, 57, 4688–4696.
18. Njuguna, J.; Michałowski, S.; Pielichowski, K.; Kayvantash, K.; Walton, A. C. Fabrication,
characterization and low-velocity impact testing of hybrid sandwich composites with
polyurethane/layered silicate foam cores. Polym Compos 2011, 32, 6–13.
19. Deng, R.; Davies, P.; Bajaj, A. K. Flexible polyurethane foam modelling and identification of
viscoelastic parameters for automotive seating applications. J Sound Vib 2003, 262, 391–417.

22
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
20. Fang, M.; Webster, T. F.; Gooden, D.; Cooper, E. M.; McClean, M. D.; Carignan, C.; Makey,
C.; Stapleton, H. M. Investigating a Novel Flame Retardant Known as V6: Measurements in
Baby Products, House Dust, and Car Dust. Environ Sci Technol 2013, 47, 4449–4454.
21. Morgan A. B. A Review of Transition Metal-Based Flame Retardants: Transition Metal Oxide/
Salts, and Complexes. In Fire and Polymers V; American Chemical Society: Washington, DC,
2009.
22. Bann, B.; Miller, S. A. Melamine and Derivatives of Melamine. Chem Rev 1958, 58, 131–172.
23. Weil E. D., Levchik S. V. Phosphorus Flame Retardants. In Kirk‐Othmer Encyclopedia of
Chemical Technology; John Wiley & Sons: Hoboken, NJ, 2017.
24. Wang, X.; Kalali, E. N.; Wan, J-T; Wang, D-Y Carbon-family materials for flame retardant
polymeric materials. Prog Polym Sci 2017, 69, 22–46.
25. Gupta, G.; Birbilis, N.; Cook, A. B.; Khanna, A. S. Polyaniline-lignosulfonate/epoxy coating
for corrosion protection of AA2024-T3. Corros Sci 2013, 67, 256–267.
26. Cui, J.; Xu, J.; Li, J.; Qiu, H.; Zheng, S.; Yang, J. A crosslinkable graphene oxide in waterborne
polyurethane anticorrosive coatings: Experiments and simulation. Compos Part B Eng 2020,
188, 107889.
27. Zhang, S.; Zhang, D.; Li, Z.; Yang, Y.; Sun, M.; Kong, Z.; Wang, Y.; Bai, H.; Dong, W.
Polydopamine functional reduced graphene oxide for enhanced mechanical and electrical
properties of waterborne polyurethane nanocomposites. J Coatings Technol Res 2018, 15,
1333–1341.
28. Karmakar, H. S.; Arukula, R.; Thota, A.; Narayan, R.; Rao, C. R. K. Polyaniline-grafted
polyurethane coatings for corrosion protection of mild steel surfaces. J Appl Polym Sci 2018,
135, 45806.
29. Wen, J. G.; Geng, W.; Geng, H. Z.; Zhao, H.; Jing, L. C.; Yuan, X. T.; Tian, Y.; Wang, T.;
Ning, Y. J.; Wu, L. Improvement of corrosion resistance of waterborne polyurethane coatings
by covalent and noncovalent grafted graphene oxide nanosheets. ACS Omega 2019, 4,
20265–20274.
30. Thakur, S.; Karak, N. Castor oil-based hyperbranched polyurethanes as advanced surface
coating materials. Prog Org Coat 2013, 76, 157.
31. Zhang, M.; Zhao, F.; Luo, Y. Self-Healing Mechanism of Microcracks on Waterborne
Polyurethane with Tunable Disulfide Bond Contents. ACS Omega 2019, 4, 1703–1714.
32. Fang, Y.; Du, X.; Jiang, Y.; Du, Z.; Pan, P.; Cheng, X.; Wang, H. Thermal-Driven Self-Healing
and Recyclable Waterborne Polyurethane Films Based on Reversible Covalent Interaction. ACS
Sustain Chem Eng 2018, 6, 14490–14500.
33. Domanska, A.; Boczkowska, A. Biodegradable polyurethanes from crystalline prepolymers.
Polym Degrad Stab 2014, 108, 175–181.
34. Li, Y-J; Nakamura, N.; Wang, Y-F; Kodama, M.; Nakaya, T. Synthesis and
Hemocompatibilities of New Segmented Polyurethanes and Poly(urethane urea)s with
Poly(butadiene) and Phosphatidylcholine Analogues in the Main Chains and Long-Chain
Alkyl Groups in the Side Chains. Chem Mater 1997, 9, 1570–1577.
35. Jaudouin, O.; Robin, J-J; Lopez-Cuesta, J-M; Perrin, D.; Imbert, C. Ionomer-based
polyurethanes: a comparative study of properties and applications. Polym Int 2012, 61,
495–510.

23
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.
36. Joshi, D. C.; Saxena, S.; Jayakannan, M. Development of l-Lysine Based Biodegradable
Polyurethanes and Their Dual-Responsive Amphiphilic Nanocarriers for Drug Delivery to
Cancer Cells. ACS Appl Polym Mater 2019, 1, 1866–1880.
37. Xinrong, S.; Nanfang, W.; Kunyang, S.; Sha, D.; Zhen, C. Synthesis and characterization of
waterborne polyurethane containing UV absorption group for finishing of cotton fabrics. J Ind
Eng Chem 2014, 20, 3228–3233.
38. Shaaban, A.; Se, S-M; Ibrahim, I. M.; Ahsan, Q. Preparation of rubber wood sawdust-based
activated carbon and its use as a filler of polyurethane matrix composites for microwave
absorption. New Carbon Mater 2015, 30, 167–175.
39. Li, X.; Yu, R.; He, Y.; Zhang, Y.; Yang, X.; Zhao, X.; Huang, W. Self-Healing Polyurethane
Elastomers Based on a Disulfide Bond by Digital Light Processing 3D Printing. ACS Macro Lett
2019, 8, 1511–1516.
40. Lei, K.; Zhu, Q.; Wang, X.; Xiao, H.; Zheng, Z. In Vitro and in Vivo Characterization of a
Foam-Like Polyurethane Bone Adhesive for Promoting Bone Tissue Growth. ACS Biomater Sci
Eng 2019, 5, 5489–5497.
41. Balcioglu, S.; Parlakpinar, H.; Vardi, N.; Denkbas, E. B.; Karaaslan, M. G.; Gulgen, S.;
Taslidere, E.; Koytepe, S.; Ates, B. Design of Xylose-Based Semisynthetic Polyurethane Tissue
Adhesives with Enhanced Bioactivity Properties. ACS Appl Mater Interfaces 2016, 8,
4456–4466.

24
Gupta and Kahol; Polyurethane Chemistry: Renewable Polyols and Isocyanates
ACS Symposium Series; American Chemical Society: Washington, DC, 2021.

You might also like