ArticlePDF Available

Synthesis of Bisphenol A Based Phosphazene-Containing Epoxy Resin with Reduced Viscosity

Authors:

Abstract and Figures

Phosphazene-containing epoxy oligomers (PEO) were synthesized by the interaction of hexachlorocyclotriphosphazene (HCP), phenol, and bisphenol A in a medium of excess of epichlorohydrin using potassium carbonate and hydroxide as HCl acceptors with the aim of obtaining a product with lower viscosity and higher phosphazene content. PEOs are mixtures of epoxycyclophosphazene (ECP) and a conventional organic epoxy resin based on bisphenol A in an amount controlled by the ratio of the initial mono- and diphenol. According to 31P NMR spectroscopy, pentasubstituted aryloxycyclotrophosphazene compounds predominate in the ECP composition. The relative content in the ECP radicals of mono- and diphenol was determined by the MALDI-TOF mass spectrometry method. The organic epoxy fraction, according to gas chromatograpy-mass spectrometry (GC-MS), contains 50–70 wt % diglycidyl ether of bisphenol A. PEO resins obtained in the present work have reduced viscosity when compared to other known phosphazene-containging epoxy resins while phosphazene content is still about 50 wt %. Resins with an epoxy number within 12–17 wt %, are cured by conventional curing agents to form compositions with flame-retardant properties, while other characteristics of these compositions are at the level of conventional epoxy materials.
Content may be subject to copyright.
Polymers 2019, 11, 1914; doi:10.3390/polym11121914 www.mdpi.com/journal/polymers
Article
Synthesis of Bisphenol A Based Phosphazene-
Containing Epoxy Resin with Reduced Viscosity
Vyacheslav V. Kireev
1
, Yulya V. Bilichenko
1
, Roman S. Borisov
2,3
, Jianxin Mu
4
,
Dmitry A. Kuznetsov
5
, Anastasiya V. Eroshenko
1
, Sergey N. Filatov
1
and Igor S. Sirotin
1,
*
1
Department of Plastics, Mendeleev University of Chemical Technology of Russia, Miusskaya sq. 9,
125047 Moscow, Russia; kireev@muctr.ru (V.V.K.); julyab2@gmail.com (Y.V.B.);
eroshenko.nast@yandex.ru (A.V.E.); filatovsn@muctr.ru (S.N.F.)
2
Topchiev Institute of Petrochemical Synthesis, Russian Academy of Sciences, Leninskii pr. 29,
119991 Moscow, Russia; borisov@ips.ac.ru
3
Department of Organic Chemistry, Peoples Friendship University of Russia, Miklukho-Maklaya str.6,
117198 Moscow, Russia
4
College of Chemistry, Jilin University, 2699 Qianjin Street, Changchun 130012, China; emujianxin@163.com
5
Scientific & Research Institute of Natural Gases and Gas Technologies — Gazprom VNIIGAZ, Razvilka,
s.p. Razvilkovskoe, Leninsky dist., Moscow region, 142717 Moscow, Russia; expdk@yandex.ru
* Correspondence: isirotin@muctr.ru; Tel.: +7-(499)-978-9265
Received: 3 October 2019; Accepted: 16 November 2019; Published: 20 November 2019
Abstract: Phosphazene-containing epoxy oligomers (PEO) were synthesized by the interaction of
hexachlorocyclotriphosphazene (HCP), phenol, and bisphenol A in a medium of excess of
epichlorohydrin using potassium carbonate and hydroxide as HCl acceptors with the aim of
obtaining a product with lower viscosity and higher phosphazene content. PEOs are mixtures of
epoxycyclophosphazene (ECP) and a conventional organic epoxy resin based on bisphenol A in an
amount controlled by the ratio of the initial mono- and diphenol. According to
31
P NMR
spectroscopy, pentasubstituted aryloxycyclotrophosphazene compounds predominate in the ECP
composition. The relative content in the ECP radicals of mono- and diphenol was determined by
the MALDI-TOF mass spectrometry method. The organic epoxy fraction, according to gas
chromatograpy-mass spectrometry (GC-MS), contains 50–70 wt % diglycidyl ether of bisphenol A.
PEO resins obtained in the present work have reduced viscosity when compared to other known
phosphazene-containging epoxy resins while phosphazene content is still about 50 wt %. Resins
with an epoxy number within 12–17 wt %, are cured by conventional curing agents to form
compositions with flame-retardant properties, while other characteristics of these compositions are
at the level of conventional epoxy materials.
Keywords: epoxy resin, epoxy oligomer, phosphazene, bisphenol A, phenol
1. Introduction
Out of all the varieties of synthetic polymers, epoxies, developed over 70 years ago, won a special
place in industry and everyday life, not necessarily in terms of overall manufacturing volume, moreso
in their specific role [1]. Classic bisphenol A based epoxy resins have many advantages: low cost, low
curing shrinkage, good chemical resistance, and mechanical properties [1,2]. In terms of valuable
qualities, epoxy polymers are superior to many other classes of synthetic polymers, which makes
them indispensable as a basis for adhesives, paints, coatings, and binders for reinforced plastics [1,3].
The development and application of new epoxy oligomers and binders based on them are expanding
at a rapid pace despite the appearance of new generation binders, such as bismaleimides, polyimides,
and cyanate esters [2]. The latter are characterized by enhanced physical and mechanical properties,
including high heat resistance and low flammability. However, their price is very high, their
Polymers 2019, 11, 1914 2 of 16
processing properties are worse than those of the same epoxides, and a high degree of crosslinking
can adversely affect the mechanical properties [1,2].
However, polymers based on unmodified epoxy resins often have low and unstable
performance characteristics, in particular flammability and relatively low heat resistance [3].
A rather effective way to increase the heat resistance of epoxy matrices is with structural
modification, for example, with compatible oligomers of higher functionality, which are incorporated
into the three-dimensional network structure formed during curing. Examples of such resins are
triglycidyl-p-aminophenol (TGPAP), tetraglycidyl-4,4-methylenedianiline (TGMDA), and
epoxidized novolacs [3,4].
The solution of the flammability problems of epoxy polymers is a more complex issue and
involves trade-offs. Thus, a traditional and effective way to reduce the flammability of epoxy systems
is to replace bisphenol A based epoxy resins with their brominated analogs. However, brominated
epoxy resins, as well as other halogen-containing compounds, emit toxic gases when they come into
contact with flames, which limits their use, for example, in civil aviation and transport, for
environmental reasons [5,6].
One of the promising ways of modifying epoxy polymers to increase heat and fire resistance is
the introduction of organometallic compounds, in particular phosphorus compounds, known as
universal flame retardants for a large number of polymeric materials. Refusal of halogen-containing
flame retardants in favor of phosphorus-based ones is a modern global trend [5].
However, the introduction of low-molecular, non-reactive compounds into the material,
including red phosphorus and its inorganic compounds (phosphates, polyphosphates, etc.) worsens
the mechanical properties and transparency of polymers [5]. Organic phosphates, which are better
combined with the base polymer, have proven themselves more functional. Meanwhile, there are
some difficulties associated, for example, with the fact that ethers of phosphoric acids can act as
plasticizers, reducing the heat resistance of the material [7]. The best properties are possessed by
binders containing functional organic phosphorus compounds capable of forming covalent bonds
with an epoxy matrix. [6,8]. For example, glycidyl ethers of phosphorus acids, which not only reduce
the flammability, but also increase the mechanical and adhesive strength of the material, are quite
promising structural phosphorus-containing modifiers [8,9]. However, they are also poorly
compatible with epoxy polymers [9,10]. Finally, most industrial phosphorus-containing flame
retardants have a much lower degradation temperature than epoxy polymers [5,6,11,12], which
makes it impossible to use them in engineering plastics and as part of high-temperature binders.
The combination of the above factors leads to the fact that epoxy resins are not always able to
satisfy the growing needs of high-tech industries, especially in the aerospace industry, automotive
industry, electrical engineering, electronics, etc.
A possible way to solve the problems described above, including the flammability of epoxy
resins, is the use of modifiers based on phosphazenes.
The main chain of organophosphazenes consists of alternating atoms of phosphorus and
nitrogen, and at the phosphorus atom there are organic radicals introduced by the substitution of
halogen in halogenphosphazenes. The nature of organic substituents, usually introduced by the
reaction of nucleophilic substitution of chlorine, can vary widely and determines the properties of
the final polymer or oligomer. The unique properties of various organophosphazenes cause the ever-
growing interest of researchers in phosphazene chemistry [13]. Compared to other
organophosphorus compounds, aryloxyphosphazenes have, as a rule, higher thermal stability and
chemical resistance and are promising non-halogen flame-retardants [6,14,15], characterized by the
synergistic action of phosphorus and nitrogen [14,16]. Thus, phenoxycyclophosphazenes were
commercialized as a flame retardant by Otsuka Chemical and others [14,17].
Researchers have long been trying to combine the exceptional properties of
organophosphazenes as highly effective flame retardants with the function of structural modifiers,
such as polyfunctional epoxy resins. There are two main synthetic approaches that allow to obtain
functional phosphazenes capable of forming covalent bonds with epoxy matrices:
Polymers 2019, 11, 1914 3 of 16
1) The synthesis of organophosphazenes with reactive epoxy groups for addition to the epoxy
component [18–43];
2) The synthesis of organophosphazenes with reactive amine groups for use as a curing agent or
its component [44–52].
Functional epoxyphosphazenes are highly effective flame retardants that not only do not reduce
mechanical properties, since they are well compatible with the epoxy matrix, but can also improve
them, probably due to the formation of a special three-dimensional polymer network, in the nodes of
which phosphazene cycles are located [53].
Currently, the majority of methods for the synthesis of functional epoxyphosphazenes described
in the literature are of primarily scientific interest due to the complexity of scaling and the large
number of intermediate stages [18–22,24,28,30,31,33–36,38–43]. Although there are
epoxyphosphazenes that are fairly easy to synthesize, for example, on the basis of hexachlorocyclo
triphosphazene and glycidol [23,25–27,29,32,37,40]. However, alkoxyphosphazenes, which include
such glycidyloxyphosphazenes, are not thermally stable [13]. Thus, even during their synthesis, an
undesirable phosphazene-phosphazane rearrangement occurs [27]. Thus, only aromatic
organophosphazenes can be used as a component of high-temperature epoxy binders.
In recent years, phosphazene-containing epoxy oligomers (PEO) with reduced flammability on
the base of cyclic chlorophosphazenes have been synthesized and characterized [54]. The most
accessible and promising are PEOs obtained by the reaction of epichlorohydrin with
hydroxyaryloxycyclotriphosphazenes (HAP), the condensation products of
hexachlorocyclotriphosphazene (HCP) and 4,4-dioxydiphenyl-2,2-propane (Figure 1).
Figure 1. Synthesis of phosphazene-containing epoxy oligomers by the reaction of epichlorohydrin
with hydroxyaryloxycyclotriphosphazenes, the condensation products of
hexachlorocyclotriphosphazene and 4,4-dioxydiphenyl-2,2-propane [55].
Hydroxyaryloxycyclotriphosphazenes (HAP), epoxycyclophosphazene (ECP), diglycidyl ether of
bisphenol A (DGEBA), phosphazene-containing epoxy oligomers (PEO).
The main problem in the synthesis of PEO is the high functionality of HCP, which requires the
use of a more than 10-fold molar excess of diphenol to achieve a more complete replacement of the
chlorine atoms in it and to avoid gelation. As for the industrial use of epoxyphosphazenes, the main
limiting factor complicating the preparation of the formulation and worsening its processing
Polymers 2019, 11, 1914 4 of 16
properties is that most epoxyphosphazenes are solids with a softening temperature of 80–100 ° C and
high melt viscosity. High average functionality can also influence processing properties.
Epoxidation of a mixture of HAP and excess of diphenol with epichlorohydrin produces PEO,
which contains fractions of the usual bisphenol A based epoxy resin and epoxycyclophosphazene
(ECP) oligomers with a variable ratio [54]. In [55,56], single-stage synthesis of PEO was realized by
direct interaction of HCP and an excess of diphenol (bisphenol A [55,56] or resorcinol [57]) in the
epichlorohydrin medium as a reagent and a solvent. The resulting PEO includes ECP consisting
essentially of tetra- and pentaepoxides of the above formula with n = 4 and 5 and a conventional
organic epoxy monomer. These PEOs in comparison with pure epoxyphosphаzenes contain organic
epoxide, which is in fact an active diluent that lowers the viscosity and average functionality to an
acceptable level for further processing. However, the viscosity of such PEO concentrates at ambient
temperature is still more than 200 Pa·s, which is a fairly high value, close to the processing limit.
The content of ECP in a mixture with the organic epoxide may be increased with simultaneous
reducing of the content of residual chlorine, the functionality of HCP and its molecular weight, which
was realized in our previous work [53] by replacing part of the chlorine atoms in chlorophosphazene
with monophenol residues (Figure 2).
Figure 2. Replacement of the part of the chlorine atoms in chlorophosphazene with monophenol
residues in order to increase the content of epoxycyclophosphazenes in a mixture with the organic
epoxide [53].
According to matrix-assisted laser desorption/ionization time-of-flight MALDI-TOF mass
spectrometry, the main components of the oligomer formed are phosphazene-containing di-, tri- and
tetraepoxides, the ratio of which in the reaction mixture can be varied by alternation of the value of
n. The average content of epoxy groups in these PEO is 13%–14% by weight. Compositions based on
diglycidyl ether of bisphenol A (DGEBA) and methyltetrahydrophthalic anhydride, containing 15%
of the obtained oligomers in the epoxy component, compared to pure DGEBA, were characterized
by more than 30% higher glass transition temperature and flexural strength, and at 75% content of
PEO were non-combustible and had limiting oxygen index (LOI) of 30 [53]. However, the need to
obtain and use of phenolates complicates the scalability of the described method.
Thus, the development of new scalable methods for obtaining epoxyphosphazene-containing
epoxy resins based on available starting materials with reduced viscosity and improved technological
properties is of great scientific and practical interest and may contribute to accelerating the
widespread use of phosphazene-containing resins as components of high-tech flame-retardant
polymer composite materials.
In order to exclude phenolate synthesis in the known method [53] (Figure 2) and to obtain pre-
formulated PEOs with lower viscosity and greater phosphorus content in the present work, by
analogy with [55,56], PEO was synthesized by direct interaction of HCP, phenol, bisphenol A, and
epichlorohydrin.
2. Materials and Methods
In the present work PEO was synthesized by direct interaction of HCP, phenol and bisphenol A
(BPA) in epichlorohydrin (ECH) medium as reagent and solvent in the presence of KOH with a role
of both HCl acceptor and epoxy-forming reagent (Figure 3, Method A).
Polymers 2019, 11, 1914 5 of 16
Figure 3. Single-stage synthesis of phosphazene-containing epoxy oligomers by direct interaction of
hexachlorocyclotriphosphazene (HCP), phenol and bisphenol A in epichlorohydrin medium (this
work).
For comparison, stepwise replacement of chlorine atoms in HCP was carried out first on
phenoxy groups, and then on the residues of bisphenol A in the presence of potassium carbonate as
HCl acceptor on the first stage and with KOH on the second (Figure 3, Method B).
2.1. Starting Materials
Hexachlorocyclophosphazene—a white crystalline substance with m.p. of 113 °C; nuclear
magnetic resonance (NMR) 31P-singlet spectrum with δP = 19.9 ppm, was obtained by the method
[58]. Potassium hydroxide in the form of 90.0% pure white pellets (JSC KAUSTIK, Volgograd, Russia)
was used without purification, the content of crystallization water determined by acid-base titration
was about 10%. Epichlorohydrin (Solvay, Tavaux, France) with the content of the main substance of
99.8% was distilled before use as a colorless liquid, b.p. 118 °C. Potassium carbonate (Sigma-Aldrich,
St. Louis, MI, USA) was dried in vacuum at 100 °C before use as a white crystalline substance in the
form of powder, soluble in water. Bisphenol A (PJSC Kazanorgsintez, Kazan, Russia) was purified
via repeated recrystallization from chlorobenzene to yield a product with m.p. of 156.5 °С. Phenol
(Sigma-Aldrich, St. Louis, MI, USA) was distilled before use into white crystals, m.p. 40.0 °C. Solvents
were purified according to known methods, their physical characteristics corresponded to the
literature data [59].
2.2. Synthesis of Epoxyphosphazenes
2.2.1. Single-Stage Synthesis of Phosphazene-Containing Epoxy Oligomers (Method A)
A 100 mL three-necked flask, equipped with a reflux condenser, a mechanical stirrer, and a
thermometer, was charged with 1 g (0.0028 mol) of HCP, 2.68 g (0.0144 mol) of bisphenol A, 0.54 g
(0.0058 mol) of phenol, and 65 mL of epichlorohydrin. The reaction mixture was heated to 50–55 °C
and thermostated at this temperature for 30 min until all solids were completely dissolved. After that,
Polymers 2019, 11, 1914 6 of 16
2.32 g (0.0414 mol) of potassium hydroxide was added and the process was conducted for 2 h at a
temperature of 60 °C. At the end of the synthesis, the hot solution was filtered off and excess solvent
was distilled off. The resulting mixture of epoxy oligomers was dried in vacuo at 85 °C. The reaction
product is a slightly colored viscous liquid. The yield was 3.96 g (71%).
2.2.2. Stepwise Synthesis of Phosphazene-Containing Epoxy Oligomers (Method B)
A 100 mL three-necked flask, equipped with a reflux condenser, a mechanical stirrer, and a
thermometer, was charged with 1 g (0.0028 mol) of HCP, 0.54 g (0.0058 mol) of phenol, 0.95 g (0.0069
mol) potassium carbonate, and 65 mL of epichlorohydrin. The reaction mixture was heated to 60 °C
and the process was conducted for 2 h at this temperature. The reaction mixture was then cooled to
50 °C, 3.28 g (0.0144 mol) of bisphenol A was charged and incubated at this temperature for 15
minutes until the bisphenol A was completely dissolved. After that, 2.32 g (0.0414 mol) of potassium
hydroxide was added and the process was conducted for 3 h at a temperature of 60 °C. At the end of
the synthesis, the hot solution was filtered off and excess solvent was distilled off. The resulting
mixture of epoxy oligomers was dried in vacuo at 85 °C. The reaction product was a slightly colored
viscous liquid. The yield was 4.18 g (75%).
2.3. Methods of Analysis
The 31P and 1H NMR spectra were measured in chloroform-d solutions with a Bruker AV-400
spectrometer (Bruker Corporation, Bremen, Germany) operating at 162 and 400 MHz, respectively.
The signals due to the deuterated solvents were used as internal references. The chemical shifts of the
signals were calculated relative to the signals of tetramethylsilane (1H) and phosphoric acid (31P),
which were used as references. The spectra were processed with the help of the MestReNova Lab
software package (Version 12.0.4, MESTRELAB RESEARCH, S.L, Santiago de Compostela, Spain).
MALDI-TOF mass spectrometric analysis was carried out on the Bruker Auto Flex II instrument
(Bruker Corporation, Bremen, Germany).
The gas chromatography-mass-spectrometry (GC-MS) study was carried out on a VARIAN-3800
CP/4000 MS chromatograph-mass spectrometer (manufactured by Varian, Palo Alto, CA, USA).
Separation of the substances was carried out using a VF-5ms chromatographic capillary column with
a length of 30 m, an internal diameter of 0.25 mm, and a fixed-phase layer thickness of 0.25 μm. The
carrier gas was helium. The injector temperature was 280 °C. The volume of the injected sample was
1 μL. The temperature of the column thermostat changed from 50 °C (1 min) to 280 °C (10 min) at a
rate of 10 °C/min. The temperature of the ion source and the interface was 230 and 280 °C. The mass
spectra of the substances were registered in the electron impact regime in the mass range 33–1000 Da.
The processing of the obtained data was carried out with the usage of computer software "Varian MS
Workstation", the identification of the substances was carried out using the NIST 11 mass-spectra
library.
The gel-permeation chromatography (GPC) was carried out on Shimadzu LC-20 Prominence
(Kyoto, Japan) chromatograph equipped with refractometric and UV detector (a wavelength of 264
nm,) and a PSS column (SDV; 300 mm × 8 mm; 1000 A, separation within 100–60,000 Da),
tetrahydrofuran (THF) was used as an eluent (1 mL/min). Molecular mass values were estimated
with the use of a polystyrene calibration curve. Elemental analysis was carried out by
spectrophotometry.
In the present paper, epoxy group content (in wt %) is the weight of oxirane groups (–CHCH2O)
divided by the total molecular weight of epoxy resin. Thus, the theoretical content of epoxy groups
was calculated by the formula:
𝐸 =
43 × 𝑛
𝑀100% (1)
where 43 and M are the molecular weights of the oxirane group and the whole epoxy resin molecule,
respectively, and n is the number of oxirane groups in the epoxy resin molecule. The experimental
epoxy group content was determined by the method of reverse acid-base titration [60–62]. In these
Polymers 2019, 11, 1914 7 of 16
methods, glycidyl groups are converted to chlorohydrin groups by dissolving the sample of epoxy
resin in hydrochloric acid acetone solution followed by titration of the excess of hydrochloric acid
with NaOH solution. The experimental epoxy group content is calculated by the formula:
𝐸 =𝑉−𝑉
×4𝑁
𝑔× 1000 100% (2)
where V is the amount of NaOH solution for titration of a sample containing epoxy resin in mL; V0 is
the amount of NaOH solution for titration of blank sample without epoxy resin in mL; 43 is the
molecular weight of the oxirane group; N is the concentration of NaOH solution, mol·L1; and g is the
mass of epoxy resin sample in g.
The viscosity of epoxy oligomers was evaluated on a Reotest-2 rotational viscometer with a
working cone-plane unit. The viscosity of the PEOs obtained in present and earlier papers was
compared with conventional low molecular weight bisphenol A based epoxy resin D.E.R.TM 332
epoxy resin (Dow Chemical, Midland, TX, USA) with the epoxy group content of 24.6%–25.1% and
epoxy equivalent of 171–175 which is referenced as DGEBA and with ERISYS® Resorcinol diglycidyl
ether RDGE (CVC Thermoset Specialties, Moorestown, NJ, USA) with the epoxy group content of
34.4%–36.4% and epoxy equivalent of 118–125.
3. Results and Discussion
The composition and structure of the phosphazene component of the resulting mixture of
epoxides were evaluated by combination of 31P NMR spectroscopy and MALDI-TOF spectrometry
(Figures 4 and 5) and the organic component by gas chromatography–mass spectrometry.
According to 31P NMR spectroscopy, irrespective of the initial mole ratio of
HCP:phenol:bisphenol A compounds with pentasubstituted triphosphazene rings predominate in
the phosphazene fraction (AB2 system with triplet δp = 20–26 ppm and doublet δp = 8–10 ppm)
together with insignificant amounts of tetraaryloxy-substituted compounds (AB2 system with
doublet at δp = 20–22 ppm and triplet at δp = 5–8 ppm).
Laser mass spectrometry data confirm the presence in the phosphazene fraction of mainly
pentaaryloxy-substituted HCP, containing residues of phenol and diphenol in various proportions
in the substituents near phosphorus atom (Figure 4, Table 1).
The content of compounds with different ratios of mono- and diphenolic aryloxy radicals as
substituents near phosphorus atom can be controlled by varying the initial ratio of
HCP:phenol:bisphenol A, as follows from Figure 4.
Figure 4. MALDI-TOF mass spectra of epoxycyclophosphazenes synthesized according to scheme A
at molar ratios of HCP:phenol:bisphenol A = 1:2:5 (a), 1:2:6 (b), 1:4 :3 (c), and 1:4:4 (d).
750 1000 1250 1500 1750 m/z
1587
1207
636
827
1017 1397
(a)
(b)
(c)
(d)
Polymers 2019, 11, 1914 8 of 16
The quantitative composition of PEOs was determined from the relative intensity of the peaks
on the MALDI-TOF spectra. On the example of the initial molar ratio of HCP:phenol:bisphenol A
equal to 1:3:5 the content of individual compounds in the phosphazene fractions formed by methods
A and B (Figure 3) is compared in Table 1.
More homogeneous is the composition of PEOs obtained according to method A (Figure 3): It
contains four basic compounds with one, two, three, and four epoxy groups (Figure 5а and Table 1).
It is noteworthy that there are no compounds with four and five epoxy groups in the B synthesis
product, but they appear in an amount up to 10 wt % of aryloxyphosphazenes with unsubstituted
OH groups of bisphenol A radicals (peaks with m/z = 960 and 1150 in Figure 5b).
Figure 5. MALDI-TOF mass spectra of epoxycyclophosphazenes synthesized according to schemes A
(a) and B (b) at molar ratios of HCP: phenol: bisphenol A = 1:3:5.
Table 1. The content of the phosphazene fractions of products synthesized by schemes A and B
according to the data of MALDI-TOF-spectrometry. The molar ratio of HCP: phenol: bisphenol A =
1:3:5.
m/z Compound Formula
1
Relative Content of the
Compound (% weight) in
the Product, Obtained via
Method A Method B
636 P
3
N
3
Cl(OPh)
5
12.3 3.6
654 P
3
N
3
Cl
4
(OPh)(OArOGly) - 3.0
768 P
3
N
3
Cl
2
(OPh)
3
(OArOGly) 1.6 7.2
826 P
3
N
3
Cl(OPh)
4
(OArOGly) 14.0 8.6
958 P
3
N
3
Cl(OPh)
3
(OArOGly)(OArOGly) - 5.0
960 P
3
N
3
Cl(OPh)
3
(OArOH)(OArOGly) - 8.5
1016 P
3
N
3
Cl(OPh)
3
(OArOGly)
2
22.2 9.2
1150 P
3
N
3
Cl(OPh)
2
(OArOH)(OArOGly)
2
- 3.0
1206 P
3
N
3
Cl(OPh)
2
(OArOGly)
3
25.0 19.3
1397 P
3
N
3
Cl(OPh)(OArOGly)
4
16.7 -
1587 P
3
N
3
Cl(OArOGly)
5
1.0 -
Polymers 2019, 11, 1914 9 of 16
Since the organic part of the synthesized epoxy-oligomers could not be evaluated well by
MALDI-TOF, a gas chromatography-mass spectrometry method was used for its analysis (Figures 6
and 7). As follows from Figure 6, the organic epoxide includes several compounds with different
retention times in a chromatographic column.
There is insufficient information in the literature about the analysis of epoxy monomers by gas
chromatography-mass spectrometry. However, given the fact that in order to achieve a satisfactory
separation and to ensure the release of substances from the column, it was necessary to increase the
temperature of the latter up to 280 °C, it can be assumed that fragmentation of the product
components could occur not only as a result of electron impact during ionization, but also as a result
of thermal exposure in a column.
The molecular weights of starting compounds and their derivatives formed as a result thermal
exposure and electron impact estimated from the mass spectra allowed us to propose their chemical
formulas which are listed in Table 2.
As follows from the data of gas chromatography-mass spectrometry, the main compounds in
the organic part of the epoxide synthesized according to both schemes are mono- and diglycidyl
ethers of bisphenol A with a predominant content of the latter (~70% and ~20% for methods A and B,
respectively).
Phenylglycidyl ether (PGE) is present in a small amount only in the product synthesized
according to method B (5%–10%). This indicates the preferential interaction of phenol not with
epichlorohydrin, but with HCP at the initial stages of the process.
The results of the analysis of the electron impact mass spectra of the fractions (Figure 6) isolated
from the chromatographic column allow us to draw the following conclusions (Table 2). The reaction
products eluted from the column as a result of electron impact upon registration of the mass spectra
may undergo some transformations, and the main among them is the transformation of the
isopropylidene group into ethylidene:
As a result, the majority of compounds fixed on mass spectra (peaks 3–8 in Figure 6) have a
molecular weight of 16 units less than expected, which in our opinion is due to the elimination of
CH4.
A similar decomposition of the isopropylidene group with the formation of a double bond was
observed earlier under conditions of the synthesis bisphenol A based
hydroxyaryloxycyclotriphosphazenes at 170 °C [60]:
The main component of the organic fraction of the resulting epoxides is diglycidyl ether of
bisphenol A (DGEBA), the highest yield (~70%) of which is achieved in the synthesis of method A
with the simultaneous feeding of starting reagents.
The compounds' formulas given in the sixth column of Table 2 show the elimination of methane
from the corresponding compounds of the formulas given in the third column of this table.
H3CCH3CH2
+CH
4
Polymers 2019, 11, 1914 10 of 16
Figure 6. Chromatograms of organic epoxides obtained by schemes A (a) and B (b).
Figure 7. Mass spectra of fractions of organic epoxides synthesized according to methods A (Figure
3) and B (Figure 4). Spectrum numbers (1–9) correspond to the numbers of fractions on
chromatograms, Figure 6.
Elution time, min
8 101214161820222426286
(1)
(3)
(4)
(5)
(7)
(8)
(9)
(a)
Elution time, min
8 101214161820222426286
(1)
(3)
(4)
(5)
(6)
(7)
(8)
(9)
(2)
(b)
(1) (2) (3) (4) (5)
(6) (7) (8) (9)
50 200 350
0
50
100
118 212
305
319
50 150 250 350 450
0
50
100
118 349
397
411
306
50 150 250 350
0
50
100
119 212 270
361
375
50 150 250 350
0
50
100
119 213 270
325
340
50 190
0
50
100
66
93
149
50 490
0
50
100
119
212
50 170 290
0
50
100
119
213
268
283
100
50 200
0
50
93
185
100
50 220 390
0
50
119
250
265
m/z m/z m/z m/z
m/z m/z m/z m/z
m/z
Polymers 2019, 11, 1914 11 of 16
Table 2. Results of chromatography-mass spectrometric analysis of organic fractions of reaction
products obtained according to schemes A and B. Molar ratio of HCP:phenol:bisphenol A = 1:3:5.
Chromatography Data (Figure 6) Mass-Spectrometric Data (Figure 7)
Peak No
Elution Time (min)
Probable Formula1 of
the Compound at the
Output of the
Column
Calculated
molecular Weight
Content2 of
the
Compound
(wt%)
The Most Likely
Formula
Calculated
molecular Weight
Observed m/z
Values of the
Products
Obtained by
Method
A B
1 7.8 PhOH 93 2.0/7.6 PhOH 93 93 93
7.9 PhOCH2CH=CH2 134 0.5/ - PhOCH2CH=CH2 134 133 none
2 9.3 PhOGly 150 0.9/4.8 PhOGly 150 none 149
3 14.6 HOArOH 228 1.0/6.3 HOArOH 212 212 212
4 15.8 HOArOGly 284 0.8/1.9 HOArOCH2CH=CH2 252 250 250
5 16.7 HOArOGly 284 2.0/8.2 HOArOGly 268 268 + 283 268 + 283
6 18.7 HOArOGly 306.5 - /4.1 HOArOGly 306.5 none 305
7 19.8 GlyOArOGly 340 66.8/45.9 GlyOArOGly 325 325 + 340 325 + 340
8 22.4 GlyOArOGly 361 21.2/2.,1 GlyOArOGly 361 361 361
9 26.8 GlyOArOGly 397 4.6/ - GlyOArOGly 397 397 none
1 Ar’ = ;
2 in the numerator for method A and in the denominator for method B.
It is noteworthy that in the reaction mixture there is practically no monoglycidyl ether of
bisphenol A (MGEBA) with a molecular weight of 284, which corresponds to peaks with an
insignificant intensity in the mass spectra of fraction 5. The main compound in fraction 5 for both
synthetic schemes is the compound with a molecular weight of 268—monoglycidyl ether of 4,4-
dihydroxydiphenol-1,1-ethenyl.
As for compounds with one chlorohydrin group and the central ethenyl group (peak 8 on the
chromatogram), the lack of a tendency to their dehydrochlorination is apparent due to both the
heterogeneity of the process (KOH is insoluble in the reaction medium) and its insignificant duration
(2 h).
From Figures 6 and 7, it follows that the organic epoxide formed according to Method A (Figure
3) together with epoxycyclophosphazene is more pure—it contains about 90% mono- and diglycidyl
ethers with minimal amounts of initial phenol and bisphenol A.
The properties of different PEO obtained in this work and other PEOs [55–57] that can be
synthesized using the one-pot method are summarized at Table 3. GPC was used (Figure 8) to
determine the relative content of phosphazene and organic fractions (Figure 8a, Table 3). For
comparison, the fractional composition of epoxyphosphazenes based on resorcinol obtained in a
previous study [57] was also determined (Figure 8b, Table 3).
Polymers 2019, 11, 1914 12 of 16
Figure 8. Gel-permeation chromatograms of PEO obtained in this by scheme A (a) with molar ratio
of HCP:phenol:bisphenol A = 1:2:6 (1), 1:4:4 (2), 1:3:5 (3); and of resorcinol-based PEO (b) obtained in
previous work [57] by interaction of HCP with resorcinol with molar ratio of HCP:resorcinol = 1:12
(1), 1:16 (2) and 1:24 (normalized to the height of the main peak).
The phosphazene component contents found from GPC data and calculated from the
phosphorus content value are close to each other and are about 50%–60% for phenol-bisphenol A
based product and are up to 45% for resorcinol-based product. The values of the weight average
molecular weight according to GPC exceed the values obtained by the MALDI-TOF method, which
can indicate both the limited applicability of the standard polystyrene calibration for these systems
and the fact that high molecular weight particles may not be fixed by the MALDI-TOF method due
to their limited volatility. However, the tendency of an increase in the average molecular weight with
an increase in the content of the phosphazene component is confirmed by both GPC and MALDI-
TOF, regardless of the nature of the starting reagents. It should be noted that the molecular weight
values of the low molecular weight organic fraction determined by GPC and GC-MS are close and
correspond to the theoretical assumption that its main component is diphenol diglydyl ether. From
elemental analysis and GPC data it follows that with a comparable content of undesirable chlorine,
the PEO obtained in this work is characterized by the both highest content of phosphorus (up to 5.4%)
and phosphazene component (up to 61%).
The usage of phenol as additional reagent lead to reduced average fuctionality of the product of
1.9–2.2, which is closer to commercial DGEBA or RDGE resins when compared to 2.2–2.5 for the
mixtures based on only bisphenol A or resorcinol. When compared to resorcinol-based phosphazene-
containing epoxy oligomers, the values of viscosity of PEO obtained in this work are slightly higher.
However, the viscosity of PEO obtained in this work at a ratio of HCP:phenol:bisphenol A = 1:4:4 and
industrial bisphenol A based epoxy resin (such as diglycidyl ether of bisphenol A, DGEBA) are
comparable (Table 3). At the same time, the viscosity of PEOs obtained at any ratio of
HCP:phenol:bisphenol A is significantly lower than that of bisphenol A based phosphazene-
containing epoxy resins. Thus, it can be expected that the processing properties of the PEO
synthesized in this work will be as close as possible to ordinary bisphenol A based epoxy resins. As
was shown in [53], such PEO are cured by conventional curing agents to form compositions with
reduced flammability, increased glass transition temperature, flexural strength, and modulus while
other characteristics of these compositions are at the level of commercially available epoxy materials.
(a) (b)
1
2
3
log(M. W.)
1
2
3
log(M. W.)
1.0 1.5 2. 0 2.5 3.0 3.5 4.0 4.5 5.0 1.5 2.0 2.5 3.0 3.5 4. 0 4.5
Polymers 2019, 11, 1914 13 of 16
Table 3. The properties of phosphazene-containing epoxy oligomers1.
Raw Reagents Ratio
Mixture Average
Functionality6
Content (wt%) Average Molecular Weight
Viscosity (Pas)
at the
Temperature of
С)
Epoxy Group
P Cl
Phosphazene
Fraction2
Entire Mixture3
Mw /Mn
Organic
Fraction
Phosphazene
Fraction 4
20 40 70
DGEBA
- 2.0 24.6–25.1 - - - 346/- - - 5.83 0.86 0.06
RDGE
- 2.0 34.4–36.4 - - - 243/- - - 1.10 0.11 0.03
HCP:BPA PEOs obtained by interaction of HCP with BPA and ECH [55,56,61]
1:8 2.5 17.1 3.1 2.7 49/49 1487/717 3405 1473 - 220 3
1:12 2.3 20.0 1.8 1.5 36/30 1212/681 1486 - 130 2
1:16 2.2 21.4 1.5 1.3 30/ 25 931/627 1492 440 78 2
HCP:Resorcinol PEOs obtained by interaction of HCP with resorcinol and ECH [57]
1:12 2.4 21.0 4.0 4.4 45/43 2350/380 2203 1054 8.33 6.15 0.36
1:16 2.3 28.6 3.0 2.4 30/32 1260/260 999 2.43 1.94 0.15
1 : 24 2.2 29.6 2.0 1.9 23/21 1130/260 957 1.71 0.45 0.05
HCP:PhOH :BPA PEOs obtained by interaction of HCP with BPA, phenol and ECH7 (this work)
1:2:6 2.2 16.1 4.6 2.2 61/54 4246/288 3405 1211 64.6 13.7 0.8
1:3:5 2.0 15.5 5.0 2.3 61/51 3459/413 1058 58.6 10.6 0.8
1:4:4 1.9 14.7 5.4 2.7 57/47 3248/317 930 9.4 6.0 0.8
1 This table in bold italics shows the literature data; 2 Found by gel-permeation chromatography (GPC)
/ Phosphorus content; 3 By GPC (See supplementary data for GPC curves); 4 By MALDI-TOF; 5 By GC-
MS; 6Calculated from phosphazene fraction content found from GPC, MALDI-TOF data, assuming
that organic component’s functionality is 2; 7By method A.
4. Conclusions
The phosphazene-containing epoxy oligomers obtained in the present work have an epoxy
group content within 15–16 wt %, phosphorus content within 4.6%–5.4% and epoxyphosphazene
component content of 50%–60%. These phosphazene-containing epoxy oligomers may be cured by
conventional curing agents to form materials with reduced flammability, while other characteristics
of these compositions are at the level of commercially available epoxy materials [53]. The viscosity of
obtained epoxyphosphazene-containing resins is comparable to conventional bisphenol A based
epoxies, and is much lower in comparison to similar epoxyphosphazene resins based on bisphenol
A. Thus, the obtained epoxyphosphazene resins may be used as a component of a binder for
composite materials, adhesives, and paints.
Author Contributions: Conceptualization, V.V.K and I.S.S.; experiment design, synthesis and data analysis,
Y.V.B., I.S.S. and A.V.E.; GC-MS analysis, D.A.K; MALDI-TOF analysis, R.S.B.; resources and materials, M.J. and
Y.V.B.; writing—original draft preparation, I.S.S.; writing—review and editing, V.V.K. and S.N.F.
Funding: The research is supported by the Ministry of Science and Higher Education and of The Russian
Federation within the framework of state contract NO. 14.574.21.0171 (unique identifier RFMEFI57417X0171).
Acknowledgments: We thank K.A. Brigadnov for technical assistance and participation in the discussions; A.S.
Tupikov and I.A. Sarychev for GPC analysis of PEO.
Conflicts of Interest: The authors declare no conflict of interest.
Polymers 2019, 11, 1914 14 of 16
References
1. Biron, M. Thermosets and Composites: Material Selection, Applications, Manufacturing and Cost Analysis;
Elsevier: Amsterdam, The Netherlands, 2013; ISBN 978-1-4557-3125-1.
2. Dodiuk, H.; Goodman, S.H. Handbook of Thermoset Plastics; William Andrew: San Diego, USA, 2013; ISBN
978-1-4557-3109-1.
3. Petrie, E.M. Epoxy Adhesive Formulations; McGraw Hill Professional: New York, USA, 2005; ISBN 978-0-07-
158908-6.
4. Flick, E.W. Epoxy Resins, Curing Agents, Compounds, and Modifiers: An Industrial Guide; William Andrew: San
Diego, USA, 2012; ISBN 978-0-8155-1708-5.
5. Visakh, P.M.; Arao, Y. (Eds.) Flame Retardants: Polymer Blends, Composites and Nanocomposites, 1st ed.;
Engineering Materials; Springer International Publishing: Cham, Switzerland, 2015; ISBN 978-3-319-03466-
9.
6. Lu, S.; Hamerton, I. Recent developments in the chemistry of halogen-free flame retardant polymers. Prog.
Polym. Sci. 2002, 27, 1661–1712.
7. Rodriguez, F.; Cohen, C.; Ober, C.K.; Archer, L. Principles of Polymer Systems, Sixth Edition; CRC Press: Boca
Raton, FL, USA, 2014; ISBN 978-1-4822-2378-1.
8. Amirova, L.M. Fosforsoderzhashhie i Metallkoordinirovannye Jepoksidnye Polimernye Materialy
[Фосфорсодержащие и металлкоординированные эпоксидные полимерные материалы]. Ph.D.
Thesis, Kazan National Research Technological University, Kazan, Russia, 2004.
9. Amirova, L.M.; Andrianova, K.A. Gradient polymeric materials based on poorly compatible epoxy
oligomers. J. Appl. Polym. Sci. 2006, 102, 96–103.
10. Amirova, L.M.; Stroganov, V.F.; Sakhabieva, E.V. Optical Materials the Based on Low Flammable Epoxy
Resins. Int. J. Polym. Mater. Polym. Biomater. 2000, 47, 43–60.
11. Horrocks, A.R.; Price, D. Advances in Fire Retardant Materials; Elsevier: Amsterdam, The Netherlands, 2008;
ISBN 978-1-84569-507-1.
12. Salmeia, K.A.; Gaan, S. An overview of some recent advances in DOPO-derivatives: Chemistry and flame
retardant applications. Polym. Degrad. Stab. 2015, 113, 119–134.
13. Allcock, H. Phosphorus-Nitrogen Compounds: Cyclic, Linear, and High Polymeric Systems; Elsevier: Amsterdam,
The Netherlands, 1972; ISBN 978-0-12-050560-9.
14. Levchik, S.V.; Weil, E.D. A Review of Recent Progress in Phosphorus-based Flame Retardants. J. Fire Sci.
2006, 24, 345–364.
15. Allen, C.W. The Use of Phosphazenes as Fire Resistant Materials. J. Fire Sci. 1993, 11, 320–328.
16. Leu, T.-S.; Wang, C.-S. Synergistic effect of a phosphorus–nitrogen flame retardant on engineering plastics.
J. Appl. Polym. Sci. 2004, 92, 410–417.
17. Phosphazene|Chemical Products|Otsuka Chemical Co., Ltd. Available online:
https://www.otsukac.co.jp/en/products/flame-retardant/phosphazene/ (accessed on 23 December 2018).
18. Fantin, G.; Medici, A.; Fogagnolo, M.; Pedrini, P.; Gleria, M.; Bertani, R.; Facchin, G. Functionalization of
poly(organophosphazenes)—III. Synthesis of phosphazene materials containing carbon-carbon double
bonds and epoxide groups. Eur. Polym. J. 2012, 29, 1571–1579.
19. Allcock, H.R.; Nelson, C.J.; Coggio, W.D. Photoinitiated graft poly(organophosphazenes): Functionalized
immobilization substrates for the binding of amines, proteins, and metals. Chem. Mater. 1994, 6, 516–524.
20. Chen-Yang, Y.W.; Lee, H.F.; Yuan, C.Y. A flame-retardant phosphate and cyclotriphosphazene-containing
epoxy resin: Synthesis and properties. J. Polym. Sci. Part A Polym. Chem. 2000, 38, 972–981.
21. Bertani, R.; Boscolo-Boscoletto, A.; Dintcheva, N.; Ghedini, E.; Gleria, M.; La Mantia, F.; Pace, G.;
Pannocchia, P.; Sassi, A.; Scaffaro, R.; et al. New phosphazene-based chain extenders containing allyl and
epoxide groups. Des. Monomers Polym. 2003, 6, 245–266.
22. Scaffaro, R.; Botta, L.; La Mantia, F.P.; Magagnini, P.; Acierno, D.; Gleria, M.; Bertani, R. Effect of adding
new phosphazene compounds to poly(butylene terephthalate)/polyamide blends. I: Preliminary study in
a batch mixer. Polym. Degrad. Stab. 2005, 90, 234–243.
23. El Gouri, M.; El Bachiri, A.; Hegazi, S.E.; Rafik, M.; El Harfi, A. Thermal degradation of a reactive flame
retardant based on cyclotriphosphazene and its blend with DGEBA epoxy resin. Polym. Degrad. Stab. 2009,
94, 2101–2106.
24. Liu, R.; Wang, X. Synthesis, characterization, thermal properties and flame retardancy of a novel
nonflammable phosphazene-based epoxy resin. Polym. Degrad. Stab. 2009, 94, 617–624.
Polymers 2019, 11, 1914 15 of 16
25. El Gouri, M.; Cherkaoui, O.; Ziraoui, R.; El Harfi, A. Physico-chemical study of DGEBA epoxy resin flame
retarded with an ecological flame retardant based on cyclotriphosphazene. J. Mater. Environ. Sci. 2010, 3,
157–162.
26. El Gouri, M.; El Bachiri, A.; Hegazi, S.E.; Rafik, M.; El Harfi, A. Fireproofing amelioration of epoxy resin
material by way a reactive flame retardant based on cyclophosphazene. Phys. Chem. News 2010, 56, 128–
137.
27. El Gouri, M.; Hegazi, S.E.; Rafik, M.; El Harfi, A. Synthesis and thermal degradation of phosphazene
containing the epoxy group. Ann. Chim. Sci. Mater. 2010, 35, 27–39.
28. Liu, F.; Wei, H.; Huang, X.; Zhang, J.; Zhou, Y.; Tang, X. Preparation and Properties of Novel Inherent
Flame-Retardant Cyclotriphosphazene-Containing Epoxy Resins. J. Macromol. Sci. Part B 2010, 49, 1002–
1011.
29. El Gouri, M.; El Bachiri, A.; Hegazi, S.E.; Ziraoui, R.; Rafik, M.; El Harfi, A. A phosphazene compound
multipurpose application-Composite material precursor and reactive flame retardant for epoxy resin
materials. J. Mater. Environ. Sci. 2011, 2, 319–334.
30. Gu, X.; Huang, X.; Wei, H.; Tang, X. Synthesis of novel epoxy-group modified phosphazene-containing
nanotube and its reinforcing effect in epoxy resin. Eur. Polym. J. 2011, 47, 903–910.
31. Bai, Y.; Wang, X.; Wu, D. Novel cyclolinear cyclotriphosphazene-linked epoxy resin for halogen-free fire
resistance: Synthesis, characterization, and flammability characteristics. Ind. and Eng. Chem. Res. 2012, 51,
15064–15074.
32. El Gouri, M.; El Harfi, A. Chemical modification of hexachlorocyclotriphosphazene—Preparation of flame
retardants and ecological flame retardant polymers. J. Mater. Environ. Sci. 2012, 3, 17–33.
33. Liu, J.; Tang, J.; Wang, X.; Wu, D. Synthesis, characterization and curing properties of a novel cyclolinear
phosphazene-based epoxy resin for halogen-free flame retardancy and high performance. RSC Adv. 2012,
2, 5789–5799.
34. Sun, J.; Wang, X.; Wu, D. Novel Spirocyclic Phosphazene-Based Epoxy Resin for Halogen-Free Fire
Resistance: Synthesis, Curing Behaviors, and Flammability Characteristics. ACS Appl. Mater. Interfaces 2012,
4, 4047–4061.
35. Feng, H.; Wang, X.; Wu, D. Fabrication of spirocyclic phosphazene epoxy-based nanocomposites with
graphene via exfoliation of graphite platelets and thermal curing for enhancement of mechanical and
conductive properties. Ind. Eng. Chem. Res. 2013, 52, 10160–10171.
36. Huang, X.; Wei, W.; Wei, H.; Li, Y.; Gu, X.; Tang, X. Preparation of heat-moisture resistant epoxy resin
based on phosphazene. J. Appl. Polym. Sci. 2013, 130, 248–255.
37. El Gouri, M.; El Mansouri, A.; El Gouri, R.; Hadik, N.; Cherkaoui, O.; Outzourhit, A.; El Harfi, A. Physical
behaviour of epoxy resin material flame retarded with a reactive flame retardant based on
cyclophosphazene. J. Mater. Environ. Sci. 2014, 5, 400–4007.
38. Lu, L.; Chen, Y.; Wang, S.; Yang, S.; Dong, X. Preparation and flame retardancy of MMT pattern synergy
intumescent flame-retardant epoxy resin. Gaofenzi Cailiao Kexue Yu Gongcheng Polym. Mater. Sci. Eng. 2014,
30, 139–144.
39. Xu, G.R.; Xu, M.J.; Li, B. Synthesis and characterization of a novel epoxy resin based on
cyclotriphosphazene and its thermal degradation and flammability performance. Polym. Degrad. Stab. 2014,
109, 240–248.
40. Liu, H.; Wang, X.; Wu, D. Novel cyclotriphosphazene-based epoxy compound and its application in
halogen-free epoxy thermosetting systems: Synthesis, curing behaviors, and flame retardancy. Polym.
Degrad. Stab. 2014, 103, 96–112.
41. Lakshmikandhan, T.; Sethuraman, K.; Chandramohan, A.; Alagar, M. Development of phosphazene imine-
modified epoxy composites for low dielectric, antibacterial activity, and UV shielding applications. Polym.
Compos. 2017, 38, E24–E33.
42. Liu, H.; Wang, X.; Wu, D. Synthesis of a novel linear polyphosphazene-based epoxy resin and its
application in halogen-free flame-resistant thermosetting systems. Polym. Degrad. Stab. 2005, 118, 45–58.
43. Wu, Y.; Wang, X.; Jiang, L.; Mao, Y.; Zhao, D. Kinetics of thermal decomposition and flame retardance of
phosphazene-containing epoxy resin. Proceedings of the 10th Asia-Pacific Conference on Combustion,
Beijing, China, 2015 June 19-22.
44. Takahashi, K.; Yamamoto, T.; Itoh, S.; Harakawa, K.; Kajiwara, M. Mechanical properties of epoxy resins
cured by various amines. Zairyo J. Soc. Mater. Sci. Jpn. 1988, 37, 454–459.
Polymers 2019, 11, 1914 16 of 16
45. Yamamoto, T.; Takahashi, K.; Kon, Y.; Harakawa, K. Curing of epoxy resin with phosphazene derivatives.
Kobunshi Ronbunshu 1988, 45, 851–856.
46. Yamamoto, T.; Takahashi, K.; Kon, Y.; Kobayashi, K. Tensile Behavior and Heat Resistance of Epoxy Resin
Cured with Phosphazene Derivatives. Kobunshi Ronbunshu 1989, 46, 177–181.
47. Takahashi, K.; Ishikawa, N.; Komori, T.; Yoon, H.-S. Curing of an epoxy resin with P
3
N
3
(NH
2
)
2
(OCH
2
CF
3
)
4
and its mechanical properties. Kobunshi Ronbunshu 1990, 47, 727–734.
48. Takahashi, K.; Ishikawa, N.; Yoon, H.-S. Mechanical properties of epoxy resins cured with P3N
3
(NH
2
)
2
(OC
6
H
4
Cl)
4
. Kobunshi Ronbunshu 1990, 47, 757–762.
49. Takahashi, K.; Ishikawa, N.; Yoon, H.-S. Resistance of epoxy resins cured with trichloro-tridimethylamino-
cyclotriphosphazene against chemical substances. Zairyo J. Soc. Mater. Sci. Jpn. 1990, 39, 1001–1006.
50. Takahashi, K.; Ishikawa, N.; Kohno, T.; Yoon, H.-S. Water resistance of an epoxy resin cured with
P
3
N
3
Cl
3
(N(CH
3
)
2
)
3
and the effect of glass flake reinforcement. Zairyo J. Soc. Mater. Sci. Jpn. 1991, 40, 458–463.
51. Takahashi, K.; Nakashima, J.; Ishiguro, S. Mechanical properties of trifunctional epoxy resin with
phosphazene derivatives. Kobunshi Ronbunshu 1994, 51, 717–723.
52. Chen, Y.M.; Liao, Y.L.; Lin, J.J. Synergistic effect of silicate clay and phosphazene-oxyalkyleneamines on
thermal stability of cured epoxies. J. Colloid Interface Sci. 2010, 343, 209–216.
53. Terekhov, I.V.; Filatov, S.N.; Chistyakov, E.M.; Borisov, R.S.; Kireev, V.V. Synthesis of oligomeric
epoxycyclotriphosphazenes and their properties as reactive flame-retardants for epoxy resins. Phosphorus
Sulfur Silicon Relat. Elem. 2017, 192, 544–554.
54. Kireev, V.V.; Bilichenko, Y.V.; Borisov, R.S.; Sirotin, I.S.; Filatov, S.N. Laser Mass Spectrometry Analysis of
the Formation of Phosphazene-Containing Epoxy Oligomers. Polym. Sci. Ser. B 2018, 60, 243–262.
55. Sirotin, I.S.; Bilichenko, Y.V.; Brigadnov, K.A.; Kireev, V.V.; Prudskov, B.M.; Borisov, R.S. Single-stage
synthesis of phosphazene-containing epoxy oligomers. Polym. Sci. Ser. B 2014, 56, 471–476.
56. Brigadnov, K.A.; Bilichenko, Y.V.; Polyakov, V.A.; Borisov, R.S.; Gusev, K.I.; Rudakova, T.A.; Filatov, S.N.;
Kireev, V.V. Epoxy oligomers modified with epoxyphosphazenes. Polym. Sci. Ser. B 2016, 58, 549–555.
57. Sarychev, I.A.; Sirotin, I.S.; Borisov, R.S.; Mu, J.; Sokolskaya, I.B.; Bilichenko, J.V.; Filatov, S.N.; Kireev, V.V.
Synthesis of Resorcinol-Based Phosphazene-Containing Epoxy Oligomers. Polymers 2019, 11, 614.
58. Sirotin, I.S.; Bilichenko, Y.V.; Suraeva, O.V.; Solodukhin, A.N.; Kireev, V.V. Synthesis of oligomeric
chlorophosphazenes in the presence of ZnCl
2
. Polym. Sci. Ser. B 2013, 55, 63–68.
59. Riddick, J.A.; Bunger, W.B.; Sakano, T.K. Organic Solvents: Physical Properties and Methods of Purification;
Wiley: Hoboken, NJ, USA, 1986; ISBN 978-0-471-08467-9.
60. Sirotin, I.S.; Bilichenko, Y.V.; Brigadnov, K.A.; Kireev, V.V.; Suraeva, O.V.; Borisov, R.S. Oligomeric
hydroxy-aryloxy phosphazene based on cyclic chlorophosphazenes. Russ. J. Appl. Chem. 2013, 86, 1903–
1912.
61. Simonov-Emel’yanov, I.D.; Apeksimov, N.V.; Kochergina, L.M.; Bilichenko, Y.V.; Kireev, V.V.; Brigadnov,
K.A.; Sirotin, I.S.; Filatov, S.N. Rheological and rheokinetic properties of phosphazene-containing epoxy
oligomers. Polym. Sci. Ser. B 2016, 58, 168–172.
62. Li, S.; Guan, W.; Wu, Z.; Lu, J.; Guo, J. An Improved Method to Determine Epoxy Index of Epoxy Resins.
Polym.-Plast. Technol. Eng. 2007, 46, 901–903.
© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
... The resulting PhER consisted mainly of tetra-and penta-epoxyphosphazenes and an organic epoxy monomer, which is essentially an active diluent that reduces the viscosity and average functionality to a level acceptable for further processing, and these PhER demonstrated increased mechanical properties and heat resistance. Due to the fact that the viscosity of such resins at ambient temperature is more than 200 Pa•s, which is a value close to the processing limit, studies were carried out to obtain PhER based on bisphenol A and phenol, including the one-stage method, with the purpose of reducing viscosity and increasing the phosphorus content to reduce flammability [50,55]. ...
... In order to obtain PhER with a lower viscosity, which ensures the processability of processing, and a higher phosphorus content, for a potential reduction in flammability, in comparison with PhER described in [52][53][54][55], while maintaining their advantages in mechanical properties, PhER were synthesized by direct interaction of HCP, bisphenol F, and epichlorohydrin. To assess the impact on the physical properties of the use of bisphenol F and the content of the phosphazene fraction, the analysis of the composition of the obtained phosphazene-containing epoxy resins and tests for mechanical properties, glass transition temperature, and viscosity were carried out. ...
Article
Full-text available
Organophosphazenes are of interest due to the combination of increased mechanical and thermal properties of polymer materials obtained with their use, however, they are characterized by a complex multi-stage synthesis. Moreover, the high viscosity of phosphazene-containing epoxy resins (PhER) makes their processing difficult. To simplify the synthesis of PhER, a one-step method was developed, and bisphenol F was chosen, which also provided a decrease in viscosity. In the current study, PhER were formed by a one-stage interaction of hexachlorocyclotriphosphazene (HCP) with bisphenol F isomers and epichlorohydrin in the presence of alkali, which was a mixture of epoxycyclophosphazenes (ECPh) with a functionality from 1 to 4 according to the results of MALDI-TOF analysis. Conventional epoxy resins based on bisphenol F, also formed during the process, showed high mechanical properties and glass transition temperature, and the reactivity of the obtained resins is similar to the base epoxy resins based on bisphenols A and F. Cured PhER had higher or the same mechanical properties compared to base epoxy resins based on bisphenol A and F, and a glass transition temperature comparable to base epoxy resins based on bisphenol F: glass transition temperature (Tg) up to 174.5 °C, tensile strength up to 74.5 MPa, tensile modulus up to 2050 MPa, tensile elongation at break up to 6.22%, flexural strength up to 146.6 MPa, flexural modulus up to 3630 MPa, flexural elongation at break up to 9.15%, and Izod impact strength up to 4.01 kJ/m2. Analysis of the composition of the obtained PhER was carried out by 1H and 31P NMR spectroscopy, MALDI-TOF mass spectrometry, X-ray fluorescence elemental analysis, and contained up to 3.9% phosphorus and from 1.3% to 4.2% chlorine. The temperature profile of the viscosity of the resulting epoxy resins was determined, and the viscosity at 25 °C ranged from 20,000 to 450,000 Pa‧s, depending on the ratio of reagents. The resins studied in this work can be cured with conventional curing agents and, with a low content of the phosphazene fraction, can act as modifiers for traditional epoxy resins, being compatible with them, to increase impact strength and elasticity while maintaining the rest of the main mechanical and processing properties, and can be used as a resin component for composite materials, adhesives, and paints.
... The benzene ring is characterized at 790 cm − 1 , 691 cm − 1 , indicating an interspersed substituted structure, which is consistent with the molecular structure of MXDA [64]. The absorption peaks at 1513 cm − 1 , 1246 cm − 1 , and 1100 cm − 1 conform to the characteristics of polyetheramines [65,66]. This indicates that the slurry contains amino groups, which gradually strengthen with temperature. ...
... As shown in Fig. 2 (a), the mixture of EP has a high viscosity, which decreases significantly with the temperature increase before 80 • C. Viscosity is a manifestation of internal friction of fluids, heating increases the thermal movement energy of molecules inside the resin, and expanding resin volume increases free space between molecules, so small molecules can implement rapid movement, thus making the viscosity of resin molecules continue to decline [34,35]; while the curing reaction of resin molecules at this point to increase the viscosity of the system has not yet been shown. Nevertheless, when the temperature exceeds 80 • C, the resin curing reaction intensifies, the system cross-linking increases rapidly, and the molecular mass increases rapidly, which makes resistance to molecular chains or other kinetic units increase significantly and thus makes viscosity rise. ...
Article
Full-text available
The elevated-and-cryogenic temperature resisting composites were prepared using tetrafunctional (AG-80) and bifunctional (E-42) epoxy resin blend as the matrix, polyetheramine (D400), and m-xylylene diamine (MXDA) as hybrid curing agents, silane coupling agent (KH-560) as a toughening agent, and micron-sized silicon oxide particles as fillers. The curing kinetics of the composites was investigated by differential scanning calorimetry (DSC), the effect of reaction temperature on the resin system's viscosity, the variation of gel time with temperature, and the heat resistance of the resin system were tested by thermogravimetric analysis. The apparent activation energy and reaction level of the curing reaction was calculated, the curing process was developed, and the elevated and cryogenic mechanical properties set at different times were tested. The results showed that the resin system's viscosity decreased with increasing temperature, 40 °C was determined as the optimum operating temperature, and the heat resistance of the composite was good. The gel time test developed a strategy of pre-curing by vacuum at 40 °C and curing at 60 °C. The compressive strength of the resin system was 99.55 MPa and 159.12 MPa after placing at −196 °C and 160 °C for 4 h, respectively. This study presents a theoretical basis for the curing process and optimization of curing parameters for elevated and cryogenic resistant modified epoxy resin composite and provides a potential application for its future in-situ utilization in deep space exploration and lunar construction.
... Bisphenol A (BPA, 4,4'-(propane-2,2-diyl)diphenol) is a major monomer component of polycarbonate (Park et al., 2017(Park et al., , 2019b, polysulfone (Park et al., 2019a), and epoxy resins (Kireev et al., 2019). Human beings are constantly exposed to BPA because they are exposed to circumstances around these chemical/materials in the real life. ...
Article
Full-text available
As human beings have been consistently exposed to bisphenol A (BPA) and bisphenol S (BPS) derived from various products, the intake of BPS/BPA to humans has been extensively studied. However, using conventional biological matrices such as urine, blood, or dissected skin to detect BPS/BPA in the human body system requires longer exposure time to them, hardly defines the pollutant source of the accumulated BPS/BPA, and is often invasive. Herein, our new approach i.e. fingerprint analysis quantitatively confirms the transfer of BPS/BPA from receipts (specific pollution source) to human skin only within receipt-handling of “20 s”. When receipts (fingertip region size; ~1 cm²) containing 100-300 μg of BPS or BPA are handled, 20-40 μg fingerprint⁻¹ of BPS or BPA is transferred to human skin (fingertip). This transferred amount of BPS/BPA can still be toxic according to the toxicity test using water fleas. As a visual evidence, a fingerprint map that matches the distribution of the absorbed BPS/BPA is developed using a mass spectrometry imaging tool. This is the first study to analyze fingerprints to determine the incorporation mechanism of emerging pollutants. This study provides an efficient and non-invasive environmental forensic tool to analyze amounts and sources of hazardous substances.
... The unique features of phosphazenes are their outstanding flexible molecular design and a wide range of homologous cyclo-and linear phosphazenes applied in many different fields. A lot of functionalized epoxy resins, benzoxazines, and amine-based curing agents with phosphazene core are proposed in the literature [6,[15][16][17][18][19][20]. Cyclophosphazenes, owing to their flexible design, have the possibility of replacing chlorine atoms in hexachlorocyclotriphosphazenes with practically any substituents that in fact determine the properties of the resultant compound. ...
Article
Full-text available
This work is devoted to the influence of phosphazene modifiers with different substituents on the curing process, thermal properties and flammability of benzoxazine resin. Novel catalysts with m-toluidine substituents were introduced. The catalytic activity of studied phosphazene compounds decreased in the row: hexachlorocyclotriphosphazene (HCP) > tetra m-toluidine substituted phosphazene PN-mt (4) > hexa m-toluidine substituted phosphazene PN-mt (6) > hexaphenoxycyclotriphosphazene (HPP), where HPP is totally inactive. Two types of catalysis: basic and acid were proposed. A brief study of resulting properties of polybenzoxazines was presented. The addition of any studied modifier caused the decrease of glass transition temperature and thermal stability of polymers. The morphology of cured compositions was characterized by matrix-dispersion phase structure. All phosphazene containing polybenzoxazines demonstrated the improved flame resistance.
... We note here that 'bisphenol A' is commonly used as a precursor molecule for the preparation of commercially important polymers like polycarbonates, 36 epoxy resins, 37 and some other types of polymers. 38,39 A 'bisphenol A' derived polymeric gel has been recently reported. 40 There are no reports of small-molecule gelators derived from 'bisphenol A' to the best of our knowledge. ...
Article
Bisphenol A, a common precursor molecule used in the preparation of some polymers, was investigated as a possible scaffold for the design and synthesis of small-molecule gelators. To this end, a small library of ten molecules was synthesized by appending different n-alkyl chains and aromatic moieties via the p-hydroxyl groups in a single step. Five of these served as excellent gelators of several polar solvents. All the gelators contained n-alkyl chains, while none of the aromatic moieties imparted the gelling ability to the molecules. The minimum gelation concentrations ranged between 1–3%, with the ether-linked gelators more efficient than the ester-linked ones. The gel-to-sol transition temperatures observed for these gels were in the range of 26–54 °C. Detailed characterizations of the gels for two representative gelators were carried out using techniques like FTIR, UV-vis absorption, scanning electron microscopy, rheology, powder XRD, and contact angle measurement. The gels showed good absorption profiles for two water-soluble dyes. Based on these results, we propose that ‘bisphenol A’ can be conveniently exploited as a versatile core gelating scaffold in the design of many other small-molecule gelators.
Article
The construction of polymer/phosphazene organic-inorganic hybrids using biomass can improve the sustainability of their preparation. In addition, introducing diverse functional units to these hybrids is a potential way to broaden their downstream applications. Herein, a benzoxazine derivative (AP-f), prepared from biomasses (apigenin and furfurylamine), is used as starting materials to construct a polymer/phosphazene hybrid. Through efficient benzoxazine-isocyanide-mechanochemistry (BIC-MC), a bio-polyamide derivative (PA-af) containing diverse molecular fragments (amide, phenolic hydroxyl, and tertiary amine) is prepared by solid-state ball-milling. Using PA-af as a phenolic source, a novel polyamide (PA)/phosphazene organic-inorganic hybrid (PACP-af) is successfully constructed by condensation with hexachlorocyclotriphosphazene (HCCP). PACP-af shows selective adsorption for cationic dye methylene blue (MB). Experimental results show that the maximum adsorption capacity of MB by PACP-af reaches 598.4 mg/g, and the adsorption process followed the pseudo-second-order kinetic and Langmuir adsorption model. PACP-af is also used as an enrichment-type Pb (II) electrochemical probe with an acceptable detection range (1–100 μM) and detection limit (0.085 μM).
Article
In this report, an open-chain ether-linked polymer (PE-AF) has been prepared via one-step co-condensation polymerization of 1,4-dibromo-2,5-difluorobenzene with bisphenol AF. The new polymer has shown a good solubility in non-polar solvents, good thermal stability (up to 300 °C), high char residue with a limiting oxygen index (LOI) of 33.5, surface morphology with random fluffy-like texture, and a non-porosity nature with a surface area (SA) of 35 m²g⁻¹. Tailoring these properties has been achieved by considering post-modification on the Bromo-sites of the polymer backbone. The post-modification was attained by applying Ullmann-coupling reaction of the Bromo-sites of PE-AF and aromatic diamines in presence of a catalytic amount of copper iodide. Upon inclusion of the aromatic diamines within the polymer’s backbone, the new polymers have depicted aggregated surface morphology (e.g. spherical particles ca. 0.3–0.1 µm), pronounced porosity nature (SA up to 354 m²g⁻¹), varied range of pore size distribution, and higher thermal stability with lower char residue contents when compared with PE-AF. The successful production of the polymers was confirmed by (C, H, O, and N) elemental analysis, solid-state ¹³C-NMR, and infrared (IR) spectroscopy. The application of Ullmann-coupling raises the opportunity to create cavities and pores within the polymer’s framework. Though this modification lowered the LOI of the polymers and their flame-retardant tendency, it enhanced their opportunity for gas capture and separation applications.
Article
Full-text available
Phosphazene-containing epoxy-resorcinol oligomers (PERO) are synthesized in one stage with the direct interaction of hexachlorocyclotriphosphazene (HCP), resorcinol, and epichlorohydrin in the presence of solid NaOH. Depending on the initial ratio of HCP:resorcinol, PERO contains from 20 to 50 wt.% phosphazene component (2.0–4.8% of phosphorus) and have an epoxy group content up to 30 %. Products are characterized using 1H and 31P NMR spectroscopy, MALDI-TOF mass spectrometry, and elemental analysis. According to mass spectrometry, the phosphazene fractions of PERO include up to 30 individual compounds with a predominance of cyclotriphosphazenes with one unsubstituted chlorine atom and four or five glycidyl groups. PERO has a lower viscosity in comparison with similar resins based on bisphenol A, which can simplify their use as a binder for polymer composites, adhesives, and paints.
Book
The Fifth Edition of Principles of Polymer Systems has been completely revised and updated. The chemical engineering perspective has been retained and strengthened, and the broad applications of polymers in chemistry and materials science have been addressed. The theoretical basis for various topics has been deepened and strengthened and several new topics are addressed. These changes reflect the rapidly growing recognition by all scientists and engineers of the role polymers play in industry. Electronics and medicine are representative areas that require more than a passing knowledge of macromolecular principles. Both areas receive attention in this edition. The end-of-chapter problems in the book have been completely replaced with the new problems. A solutions manual will be available to qualified instructors.
Article
The main features of two methods for the synthesis of phosphazene-containing epoxy oligomers— namely, the methods based on oxidation of double bonds in organooxyphosphazenes and on the reaction of chlorocyclophosphazenes with diphenols and the subsequent interaction of the resulting hydroxy-aryloxy phosphazenes with epichlorohydrin—were examined. Using the example of hexa- and octa-eugenol derivatives of the corresponding cyclophosphazenes, optimal conditions were established for the oxidation of allyl groups of these compounds with peroxy acids and hexa- and octa-epoxide cyclophosphazenes were characterized. It was noted that the epoxidation of eugenol derivatives of a mixture of cyclophosphazenes with three to eight phosphazo groups is accompanied by side reactions leading to the formation of P–OH bonds and the partial opening of oxirane cycles. Bisphenol A phosphazene-containing oligoepoxides were synthesized both via the stage involving the formation of hydroxy-aryloxy cyclophosphazenes and their subsequent epoxidation with epichlorohydrin and via the direct interaction of chlorocyclophosphazenes with an excess of bisphenol A (BPA) in the presence of solid alkali. In the latter case, the resulting oligomers are mixtures of the conventional epoxide and phosphazene-containing epoxy oligomers. The content of the latter can be adjusted up to 50%. The synthesized oligomers contain 1–5% phosphorus. They can be cured by conventional hardeners to form flameproof or noncombustible compositions.
Article
Hydroxyaryloxycyclophosphazenes containing 2-4 OH groups have been synthesized by the substitution of chlorine atoms of hexachlorocyclotriphosphazene via the reaction with sodium phenolates of halogenophenols followed by the interaction with sodium monophenolate of diphenylolpropane. Oligoepoxyphosphazenes (OEPs) with molecular masses up to 2000 and the contents of epoxy groups, phosphorus, and halogens atoms about 5-8, 5-8, and 5-11%, respectively, have been obtained via the interaction of the aforementioned phosphazenes with epichlorohydrin. The curing of the OEPs with amines or acid anhydrides gives rise to the formation of self-extinguishing composites. The incorporation of the OEPs (5-75 wt %) into commercial epoxy resins followed by their curing, results in the formation of composites with excellent nonflammability or capability of self-quenching, and good dielectric, heat resistant and mechanical properties. It has been established that mixtures of common epoxides with different amounts of the OEPs can be synthesized by a "single-reactor" method.
Article
Phosphazene-containing epoxy oligomers are obtained through a one-stage method via the interaction of hexachlorocyclotriphosphazene with an excess of diphenylolpropane in the medium of epichlorohydrin in the presence of a solid alkali. With the use of ³¹Р NMR and MALDI mass spectrometry, it is shown that the main components of the phosphazene fraction are penta- and tetra-aryloxysubstituted cyclotriphosphazene compounds with the corresponding numbers of epoxy groups. The maximum content of the phosphazene fraction in phosphazene-containing epoxy oligomers (~40 wt %) is attained at a hexachlorocyclotriphosphazene-to-diphenylolpropane molar ratio of 1: 8. Phosphazene-containing epoxy oligomers cured with isomethyltetrahydrophthalic anhydride have oxygen indexes of 26–28 and are self-extinguishing.
Article
The rheological and rheokinetic properties of a two-component binder consisting of epoxy-diane oligomers and the oligoepoxyphosphazenes PEO-1 (30 wt %) and PEO-2 (40 wt %) are studied. The viscosities of the initial oligomers at 40°C are 130 (PEO-1) and 270 (PEO-2) Pa s; the activation energies of viscous flow in the range 40–70°C are from 122 to 128 kJ/mol. The addition of equivalent amounts of curing agents, such as triethylenetetramine or iso-methyltetrahydrophthalic anhydride, reduces the initial viscosity of a composition, most strongly in the presence of the second curing agent (by a factor of 50–100). The activation energies of the cure process with triethylenetetramine in the range 45–95°C are 89 (PEO-1) and 125 (PEO-2) kJ/mol, and the gelation time at 55°C is 6 min for both oligomers. The time of gelation for the system PEO–iso-methyltetrahydrophthalic anhydride at 90°C is 475 min, and the glass-transition temperatures of the cured compositions are 238 (PEO-1) and 250°C (PEO-2), as evidenced by thermomechanical studies.
Book
This book bridges the technology and business aspects of thermosets, providing a practical guide designed for engineers working in real-world industrial settings. The author explores the criteria for material selection, provides information on material properties for each family of thermosets, and discusses the various processing options for each material type. He explains advantages and disadvantages of using thermosets and composites in comparison to competing materials and assesses cost aspects, enabling the reader to balance out technical and economic constraints when choosing a thermoset and processing technology for a given application. This second edition contains a new section on composites solutions for practical problems, gathering information on trends contributing to the breakthrough of composites in various sectors. Other new sections on specific crosslinking processes, processing trends, machinery and equipment manufacturers, applications, bio-sourced thermosets and natural fibers, and recycling of thermosets and composites are included. Case studies are provided, illustrating many design and production challenges. Furthermore, new market data and information about health and safety will be added. All data is fully updated throughout, with pricing in USD and EUR, and both ASTM (North American) and European standards. Thermosets and Thermoset Composites, Second Edition is the only book that gives in-depth coverage of a wide range of subject matters and markets, yet in brevity and concision in a single volume, avoiding the need of consulting a series of other specialized books. By providing the knowledge necessary for selecting a fabrication process, thermoset material and methods for determining the all important cost of thermoset parts this new edition is an invaluable decision-making aid and reference work for practitioners in a field with growing importance.
Book
This important book provides a comprehensive account of the advances that have occurred in fire science in relation to a broad range of materials. The manufacture of fire retardant materials is an active area of research, the understanding of which can improve safety as well as the marketability of a product. The first part of the book reviews the advances that have occurred in improving the fire retardancy of specific materials, ranging from developments in phosphorus and halogen-free flame retardants to the use of nanocomposites as novel flame retardant systems. Key environmental issues are also addressed. The second group of chapters examines fire testing issues and regulations. A final group of chapters addresses the application of fire retardant materials in such areas as composites, automotive materials, military fabrics and aviation materials. With its distinguished editors and array of international contributors, this book is an essential reference for producers, manufacturers, retailers and all those wishing to improve fire retardancy in materials. It is also suitable for researchers in industry or academia.
Article
Flame retarded epoxy materials were prepared by blend of HGCP (Hexaglycidyl cyclotriphosphazene) in an organic matrix based on DGEBA epoxy resin. The blends thermoset with 4,4'-methylene-dianiline (MDA) curing agent were characterized by various techniques which consist in thermal by DSC, the morphology by SEM. In electrical characterization it's done with frequency variation range from 100Hz to 100 KHz at room temperature. The rheological behaviour was investigated using small-deformation rheology. These measurements revealed that the reological and electrical behaviours depend strongly on the HGCP amount in DGEBA matrix. The Capacitance-Frequency and conductance-Frequency measurements suggest a distribution of free volume in the blends and select the samples as dielectric materials. The study of linear viscoelastic properties shows that the storage modulus G' is very sensitive to HGCP amount in DGEBA epoxy resin. The as prepared materials seem to be promising materials for electronic compounds.
Article
An overview of recent developments of the chemistry of halogen-free cyclotriphosphazene is presented in this paper. The polymers or reactive monomers containing cyclotriphosphazene are inherently flame retarding with P and N elements. They can be used on their own or added to current bulk commercial polymers to enhance flame retardancy. The chemical modification of cyclotriphosphazene is discussed along with thermal stability and flame-retardant properties of the subsequent materials.