Sie sind auf Seite 1von 29

Polymer International Polym Int 53:19011929 (2004)

DOI: 10.1002/pi.1473
Review
Thermal decomposition, combustion and
ame-retardancy of epoxy resinsa review
of the recent literature
Sergei V Levchik
1
and Edward D Weil
2
1
Akzo Nobel Chemicals, 1 Livingstone Avenue, Dobbs Ferry, NY 10522, USA
2
Polytechnic University, Six Metrotech Center, Brooklyn, NY 11201, USA
Abstract: An overview of the recent literature on combustion and ame-retardancy of epoxy resins is
presented. A brief overview of the structures of cured epoxy resins is also presented as a background
for better understanding of the thermal decomposition and combustion phenomena. The literature
sources were mostly taken from the publications of 1995 and later; however, for basic descriptions of the
structural and thermal decomposition principles, older publications are also cited. New developments in
ame-retardant additives, epoxy monomers and curing agents are described. It is shown that the main
attention in recent years has been focused on phosphorus-containing epoxy monomers and epoxy resins.
Silicon-containing or nitrogen-containing products and inorganic additives remain of great interest as
supplementary materials to phosphorus ame-retardants.
2004 Society of Chemical Industry
Keywords: epoxy resins; thermal decomposition; combustion; toxicity of combustion gases; ame retardancy
INTRODUCTION
Epoxy resins are characterized by the presence of
epoxide groups prior to cure, and they may also con-
tain aliphatic, aromatic or heterocyclic structures in
the backbone. Epoxy resins are relatively expensive;
however, the long service time and good physi-
cal properties often help by providing a favorable
costperformance ratio when compared to other ther-
mosets. The main elds where re-retardancy of epoxy
resins is required are electronics (printed wiring boards
and semiconductor encapsulation) and transportation
(automotive, high speed trains, military and commer-
cial aircraft) in composite structural and furnishing
elements.
1
Like other thermoset resins, epoxy resins
can be rendered re-retardant either by incorporating
re-retardant additives or by copolymerization with
reactive re retardants. Additive-type re retardants
are mostly used in coating or encapsulation, whereas
reactive ame retardants are preferable in printed cir-
cuit boards and composites in order to avoid the risk
of deterioration of physical properties.
STRUCTURES OF EPOXY RESINS
In order to understand the thermal decomposition and
combustion of epoxy resins, we will briey discuss the
main structural elements of the cured epoxy network.
More detailed discussion of epoxy chemistry can be
found in the books edited by May
2
and Ellis.
3
The
structure of the cured resin depends on the epoxy
monomer and curing agent used. Usually, chemical
linkages generated by reactions of glycidyl ethers
are less stable than other chemical linkages in the
epoxy network, and therefore there is justication
to discuss here only those structures formed by the
glycidyl ethers.
Carboxylic acids easily react with epoxies; however,
generally anhydrides are used for curing (Scheme 1).
Anhydrides of dicarboxylic acids produce linear
structures with diepoxides, and crosslinking usually
occurs due to esterication of the alcohol groups
(Scheme 2).
Aliphatic and aromatic diamines are the most
widely used classes of curing agents. With the
proper catalyst, aliphatic diamines cure epoxies at
room temperature, whereas elevated temperatures are
required for aromatic diamines (Scheme 3).
Secondary amines, although more hindered than
primary amines, can still react with epoxies and form
crosslinks.
Aliphatic alcohols are reactive with epoxies; how-
ever, they are not normally used as curing agents,

Correspondence to: Sergei V Levchik, Akzo Nobel Chemicals, 1 Livingstone Avenue, Dobbs Ferry, NY 10522, USA
E-mail: sergei.levchik@akzo-nobel.com
(Received 24 February 2003; revised version received 9 May 2003; accepted 3 September 2003)
Published online 13 October 2004
2004 Society of Chemical Industry. Polym Int 09598103/2004/$30.00 1901
SV Levchik, ED Weil
Scheme 1.
Scheme 2.
Scheme 3.
whereas polyphenols, which are becoming more fre-
quently used because they favor a high glass transition
temperature of the cured epoxy (Scheme 4).
In the presence of catalyst, epoxies can undergo
self-curing. This process always occurs to some extent
when the curing agent is used, and mostly concerns
the reaction of a secondary alcohol with the epoxy
(Scheme 5).
Self-curing of epoxies is often responsible for
crosslinking when difunctional curing agents are
used.
THERMAL DECOMPOSITION
The base resin
The thermal stability of epoxy resins, as well as
their ammability, depends on the structure of the
monomer, the structure of the curing agent and the
crosslink density. The literature data on the effect of
crosslink density on thermal stability are contradictory.
For example, Dyakonov et al
4
showed that the thermal
stability increases with increasing crosslink density
for the analogous resins. On the other hand, Iji and
Kiuchi
5
found that the pyrolysis resistance of novolac
epoxy resins was slightly decreased by the addition
of an excess of hardener. In general, the thermal
stabilities of aromatic epoxy resins are higher than
those of aliphatic ones, even though the crosslink
densities of the aromatic networks may be lower. The
copolymerization of aliphatic or aromatic epoxy-amine
cured resin with self-cured resol novolac resin leads
to an increase of crosslinking and thermal stability;
however, the level of the epoxy-amine system should
be kept below 15wt%.
6
Christiansen et al
7
observed that epoxies cured with
dicyandiamide (DICY) performed relatively poorly
in thermal stability tests, while epoxies cured with
novolacs had the best thermal stability. This was
explained by the fact that the amine linkages present
in the cured epoxy exhibit a lower thermal stability
than ether linkages.
Usually, thermal decomposition of any epoxy
resin starts from the dehydration of the secondary
alcohol, leading to the formation of vinylene ethers
821
(Scheme 6).
Scheme 4.
Scheme 5.
1902 Polym Int 53:19011929 (2004)
Decomposition, combustion and re retardancy of epoxy resins
Scheme 6.
Scheme 7.
Scheme 8.
Scheme 9.
The resulting allylic ether CO bond is thermally
less stable than the original CO, and therefore
chain scission mostly occurs at the allylic position.
In epoxies copolymerized with amines, the allylic
amine CN bond is less stable than the allylic ether
CO bond, and therefore in general, the amine-
cured epoxy resins are less stable than anhydride-cured
epoxies.
22
It was found that epoxies cured with meta-
substituted aromatic amines are more thermally stable
than epoxies cured with para-substituted aromatic
amines.
4
Both homolytic and heterolytic chain scission
of allylic CN (CO) bonds were suggested in the
literature,
13,16,18,23
which leads essentially to the same
result (Scheme 7).
A second similar chain scission with the secondary
amine group (CNH) might be expected to result in
liberation of the curing agent. However, this seems not
to be the case, since amines were found only as a minor
product of the thermal decomposition of epoxies.
24
Amines usually volatilize as a part of the chain
fragments or stay in the solid residue and undergo
charring. On the other hand, phthalic anhydride was
regenerated in large quantities on thermal decomposi-
tion of anhydride-cured epoxy resins
22,25
(Scheme 8).
Upon further decomposition, aliphatic chain ends
produce light combustible gases, allyl alcohol, acetone
and various hydrocarbons.
1416,18,2022
Alternatively, the allylic ethers or amides, formed
after losing water, can undergo the Claisen rearrange-
ment (Scheme 9) which changes the paraphenylene
group to a 1,2,4-trisubstituted benzene with increased
thermal stability.
26
This structure is partially respon-
sible for further crosslinking and charring of epoxy
resins.
In addition to the main processes of chain scission
discussed above, many secondary processes, which
lead to minor products of thermal decomposition,
have been reported in the literature.
1216,18
For
example, cyclization of aliphatic chain ends, instead
of splitting off, can contribute to the charring
and re retardancy
810,1315,18,20
(Schemes 10 and
11).
At high temperatures, polyaromatic hydrocarbons,
such as naphthalene and phenanthrene, are impor-
tant degradation products formed by pyro-synthesis
and aromatization reactions.
25
Radical recombination
is also a possible pathway for the formation of highly
aromatic compounds. Biphenyl could be formed by
combination of two phenyl radicals. Hydroxydiphenyl-
methane could originate from the combination of
a phenyl radical with a hydroxyphenylmethyl radi-
cal or a phenylmethyl radical with a hydroxyphenyl
radical.
Polym Int 53:19011929 (2004) 1903
SV Levchik, ED Weil
Scheme 10.
Scheme 11.
Thermal oxidation
Thermal oxidative decomposition of the epoxy
resins has also been extensively reported in the
literature.
11,2731
Basically, three mechanisms for the
oxidation of epoxies were suggested: (1) attack of
oxygen on the methylene group,
11
(2) oxidation of
the tertiary carbons in the aliphatic portion of the
chain, which is usually an ester-type crosslink in the
anhydride cured resins,
29
and (3) oxygen attack on
the nitrogen in the amine-cured epoxies.
27,31
Any of
these mechanisms leads to the formation of carbonyl
groups (isomerization in the case of (3)) which further
decompose and result in chain splitting.
Kinetics of thermal decomposition
The kinetics of the thermal decomposition of epoxy
resins in relation to their ammability and to model
combustion has been reported in the literature.
3235
These studies applied the method of the invariant
kinetic parameter (IKP)
36
, in order to describe the
degradation of the resins. This method allows the
evaluation of the probabilities for various kinetic
functions and thus to describe the degradation
model of the material. The kinetic model which
best describes the decomposition of epoxy resins
is a diffusion-mechanism model. Comparison of the
invariant activation energies of epoxies cured with
diaminodiphenyl sulfone (DDS) and dicyandiamide
(DICY) showed that the resin cured with the aromatic
diamine is more stable towards thermal oxidation.
Similarly, Le Huy et al
30
showed that thermal
oxidative decomposition of epoxies at relatively low
temperatures is diffusion-controlled because of the
need for oxygen penetration to the deep layers of the
resin.
Decomposition in the presence of
ame-retardants
Paterson-Jones et al
15
showed that commonly used
inorganic llers, eg alumina or silica, accelerate the
thermal decomposition of epoxies, probably because
of catalysis. Kinetic evidence of the catalytic action
of Al, Cu or Zn used as llers in epoxy resins was
provided by Sanchez et al.
37
Scheme 12.
The thermal degradation of a brominated epoxy
resin consisting of the diglycidylether of bisphe-
nol A (DGEBA), chain-extended with tetrabro-
mobisphenol A (TBBA) and cured with 4,4

-
diaminodiphenylsulphone (DDS), has been investi-
gated by Luda et al.
38
Thermolysis occurred in three
steps: decomposition of the brominated part of the
resin, decomposition of the non-brominated part and
char formation. Brominated aliphatics, mono- and
dibrominated phenols were released in the rst stage,
whereas HBr was found in the gaseous products only
above 330

C. The non-brominated part of the resin


decomposed with evolution of unsubstituted and alkyl-
substituted phenols, bisphenol Aand alkoxyaromatics.
Nitrogen-containing groups were found to accumulate
in the residue due to the high level of crosslinking.
The kinetics and mechanisms of the thermal degra-
dation of a phosphorus-containing epoxy, based on
bis-(3-glycidyloxy)phenylphosphine oxide (BGPPO)
(Scheme 12) and 4,4

-diaminodiphenylsulphone
(DDS), were studied by Liu et al.
39
The degradation
of the epoxy resulted in high char yields and residues
with high phosphorus contents. The phosphorus con-
tent of the residues of thermal degradation was found
to increase from 6.2wt% at 200

C to 9.7 wt% with


increasing temperature to 550

C. It was suggested
that phosphorus plays an important role in xing the
carbon, to reduce the production of ammable gases
and to form a phosphorus-rich residue.
The thermal properties of epoxy resins contain-
ing ame-retardants based on silicon, phosphorus
and melamine, were investigated by using thermo-
gravimetric analysis.
40
Phosphorus groups lowered
the initial decomposition temperature of the epoxy
resin, whereas silicon and melamine groups did not.
The integral decomposition temperatures of the epoxy
resins were signicantly increased with simultaneous
incorporation of phosphorus and silicon, owing to
the formation of highly thermally and oxidatively sta-
ble char which led to good char preservation at high
temperatures under air. The low char yields of the
epoxy resins containing only silicon also implied that
silicon could serve as a char-protector, but not a pro-
moter for char-formation. Incorporating melamine
groups into the silicon-containing epoxy resins did
not alter the resins thermal stability and degradation
characteristics.
COMBUSTION
Epoxy resins are very reactive, which allows versatility
of curing agents, either catalytic or reactive. The
catalytic curing agents do not build themselves into
1904 Polym Int 53:19011929 (2004)
Decomposition, combustion and re retardancy of epoxy resins
the thermoset structure and therefore do not much
affect the ammability of the resin. For example, it
was shown
41
that epoxy resins catalyzed by various
boroxines have essentially the same Oxygen Index (OI)
because the crosslinking densities were comparable.
On the other hand, some enhancement of the oxygen
index was observed upon increasing the amount of the
catalyst and a very signicant increase was observed
upon increasing of curing time, which is attributed to
the increase of the crosslink density.
42,43
In contrast,
Macaione et al
44
found no essential change in the
OI once the system has been through the initial
cure portion of the cure cycle so that the presence
or absence of post-curing makes no contribution to
the ammability characteristics of the material. A
very high crosslink density of the epoxy resin could
make the network structures too rigid to produce
intumescent charred layers during combustion and
therefore deteriorate ame retardancy.
45
The chars
of highly crosslinked resins instead cracked and
continued burning.
Reactive curing agents mostly represented by
amines, anhydrides or phenolic resins strongly modify
the ammability. The combustion behavior of simi-
lar epoxy resins depends on the ratio of oxygen to
carbon atoms in the polymer structure. Epoxy resins
cured with amines tend to produce more char and they
apparently are less ammable than acid- or anhydride-
cured resins at comparable crosslink densities.
46
How-
ever, nitrogen can be supplied also from the epoxy
monomer. For example, a self-cured tetraglycidyl
diaminodiphenylmethane-type resin containing nitro-
gen in its structure is less combustible
20
than the same
resin cured by diaminodiphenyl sulfone.
21
Lin and Pearce
47
and Chen et al
48
showed that
the OI of epoxy resins correlates with the charring
performance. Le Bras et al
35
found correlations
between OI and peak of heat release rate and ignition
times measured in cone calorimetry. Typically, non-
ame-retardant resin ignites at about 4560s and
extinguishes at about 200220s at a heat ux of
35kWm
2
. After 200s, epoxy resins usually show
post-glowing combustion. The combustion of the
epoxies in the cone calorimeter is not complete and
leads to the formation of thermally stable residues.
Typically, epoxy resins are more ammable than
phenolic resins used in the same application
49
and
therefore epoxy resins require ame-retardants.
FLAME-RETARDANT EPOXIES
Inherently ame-retardant epoxies
Epoxy resins cured by phenol formaldehyde resin
are to some degree inherently ame-retardant
because of the signicant charring tendency of
the phenolic component. The char yield can be
increased by using special epoxy monomers contain-
ing highly aromatic bisphenols (eg phenolphthalein,
bisphenol-uorenone,
41,47,50
bisphenol-anthrone and
Scheme 13.
Scheme 14.
Scheme 15.
Scheme 16.
tetraphenol-anthracene
48
) or those containing dou-
ble bonds able to undergo the DielsAlder reaction
(eg mono-, di- or trihydroxystyrylpyridine).
42
These
monomers were cured either alone or in combination
with regular grade epoxy monomers (eg the com-
mercial diglycidyl ether of bisphenol A (DGEBA)),
resulting in a higher crosslink density and showing an
increase of OI from 20 for DGEBA to almost 40 for
highly aromatic monomers.
Compounds that contain novolac derivatives (phe-
nol novolac-type (Scheme 13) or o-cresol novolac-type
(Scheme 14) epoxy resins) including aromatic groups
in the main chain display far higher ame-retardancy
than that of epoxy resin compounds without aro-
matic groups.
5,45,51
Of the two aromatic groups
included in the chain, the biphenylene group (a mix-
ture of 4,4

-diglycidyl-(3,5,3

,5

tetramethyl-biphenyl)
ether (Scheme 15) and 4,4

-diglycidyl-biphenyl ether
(Scheme 16) (50 + 50wt %)) was more effective than
the phenylene group. However, the inclusion of a
non-aromatic group, a phenol dicyclopentadiene-type
epoxy resin (Scheme 17), in the chain had no ame-
retardant effect.
Epoxy resin compounds containing the biphenylene
group in the novolac structure (Scheme 18) were more
effective than compounds containing the phenylene
group (Scheme 19).
It was speculated that the biphenylene-containing
resin resulted in lower crosslinking densities and higher
Polym Int 53:19011929 (2004) 1905
SV Levchik, ED Weil
Scheme 17.
Scheme 18.
Scheme 19.
Scheme 20.
elasticity which facilitated formation of an intumes-
cent layer. Epoxy resins containing these novolac
derivatives with aromatic bridging groups showed
relatively high pyrolysis resistance while having rel-
atively low crosslink densities. The high pyrolysis
resistance appeared to have made an important con-
tribution to the stability of the intumescent foam
layers during combustion. As previously noted, the
inclusion of a non-aromatic group (dicyclopenta-
diene) in the novolac resins failed to form any
protective foam layer in spite of the low crosslink
density of the epoxy resin compound. This was
believed to be due to the lower pyrolysis resis-
tance of the compound with the non-aromatic
group.
In addition to the highly charrable biphenylene
epoxies shown above, an epoxy having a double bond
in its structure (Scheme 20), a naphthalene-based
epoxy (Scheme 21)
52
and epoxy resins containing tert-
butyl-substituted aromatics (Schemes 22 and 23)
53
were investigated.
Typically, these resins cured with biphenol-
(Scheme 24) or naphthalene- (Scheme 25) based
novolacs showed a V-1 UL 94 rating without any
Scheme 21.
Scheme 22.
Scheme 23.
Scheme 24.
Scheme 25.
Scheme 26.
auxiliary ame-retardant additive, or V-0 in the sys-
tems heavily loaded with silica.
The use of a multifunctional epoxy resin with four
glycidyloxy groups, tetrakis (glycidyloxyphenyl)ethane
(Scheme 26), in combination with other epoxy resin
compounds improved the thermal stability while
still maintaining excellent ame-retardancy.
5,54
It is
believed that this combination of results could be
attributed to a local increase in crosslink density.
Although the overall elasticity slightly decreased, it
remained high enough for intumescent layer forma-
tion. Another highly crosslinkable epoxy performed
similarly (Scheme 27).
54
1906 Polym Int 53:19011929 (2004)
Decomposition, combustion and re retardancy of epoxy resins
Scheme 27.
A high-performance epoxy-imide polymer was
obtained by curing diimide-diepoxide (Scheme 28)
with 4,4

-diaminodiphenyl ether (DDE), 4,4

-di-
aminodiphenylsulphone (DDS), tris(3-aminophenyl)
phosphine oxide
55
or bis(3-aminophenyl)methyl-
phosphine oxide.
56
The thermal and ame resistances
of epoxy resins were signicantly improved by the
introduction of imide groups into the epoxide struc-
ture.
In order to prepare high-performance epoxy resins,
Hwang and Jung
57
used three types of diamines: N,N

-
(4,4

-diphenylether)-bis(4-aminophthalimide) (Sch-
eme 29), 4,4

-bis(p-aminophenoxy)dibenzalpent-
aerythritol (Scheme 30) and 2,2

-bis [4-(p-amino-
benzoyl)phenyl]propane (Scheme 31) to cure DGEBA
epoxy resins. The char yields measured thermogravi-
metric analysis (TGA) at 500

C were 49.4, 46.2 and


34.5 wt% for these diamines, respectively, whereas
the char yield of the epoxy cured with methylene
bis-dianiline was 32.9wt%. The high char yield of
the imide-containing amine implies that incorporating
imide into the epoxy matrix moiety improves the ame
retardancy of the epoxy.
In order to provide better ame-retardant perfor-
mance and physical properties, epoxy resins can be
copolymerized with other thermoset resins. For exam-
ple, copolymerization of an aliphatic epoxy resin,
partially cured by an aliphatic polyamine, with resol
novolac phenolic resin, provided a signicantly better
ame-retardant material than the plain epoxy.
58
In the
case of a composite material made with the pure epoxy,
the char was formed but the bers were exposed at the
surface, due to delamination which was attributed to
Scheme 29.
Scheme 31.
stresses produced by trapped gases within the compos-
ite. The delamination of the composite during burning
was avoided in the presence of resol novolac. Tyberg
et al
59,60
prepared void-free networks from the reac-
tion of phenolic novolacs with various epoxides. These
systems had signicantly improved mechanical prop-
erties over commercial novolac networks and were
found to be comparable in re resistance to typical
phenolics as measured by cone calorimetry.
A new ternary system based on epoxy, benzoxazine
and phenolic resin was studied by Rimdusit and
Ishida.
61
It was found that the combined system
shows a relatively high decomposition temperature,
ie up to 370

C, compared with about 270

C for
the polybenzoxazine used. The material had improved
thermal stability with an increasing mass fraction of
epoxy in the system, which may be attributed to a
greater crosslink density. On the other hand, the char
yield of the ternary systems was signicantly higher
when compared with the pure epoxy resin. This is due
to the fact that both polybenzoxazine and phenolic
novolac are known to give a higher char yield when
compared with the epoxy resin.
A new advanced thermoset material was developed
by copolymerization of special low-molecular-weight
poly(phenylene oxide) (PPO) and epoxy resins.
62
The
low-molecular-weight PPO resin provides excellent
processability and compatibility with epoxy resins, in
contrast to the high-molecular-weight resin which is
traditionally used as a component in blends with high-
impact polystyrene. In addition to contributing good
heat resistance and electrical properties to the epoxy
resins, the low-molecular-weight PPO also improves
dimensional and hydrolytic stability, chemical resis-
tance and provides ame retardancy. The electronic
Scheme 28.
Scheme 30.
Polym Int 53:19011929 (2004) 1907
SV Levchik, ED Weil
Scheme 32.
industry is also practicing copolymerization of epoxy
resins with isocyanurate esters in order to achieve high
electrical performance and good ame-retardancy.
63
Halogenated ame retardants
Tetrabromobisphenol A (TBBA) (Scheme 32) is a
unique reactive ame-retardant additive widely used
in epoxy resins, especially in electronic grades, where
ame-retardancy is mandatory. Normally, TBBA is
pre-reacted with epoxy in the so-called chain extension
process. In order to obtain a V-0 rating in printed
wiring boards, 20 to 37wt% of TBBA is used. TBBA
is reportedly the highest-volume brominated product
sold in the market.
Although TBBA does not create obvious toxicolog-
ical problems, recently it fell under attack by being
classed together with halogenated additive products.
64
The move to replace halogenated materials is gath-
ering momentum with some Japanese and European,
particularly Scandinavian, original equipment manu-
facturers (OEMs).
65
The thermal stability of bromi-
nated epoxy resins has improved since the 1970s
by proper formulation with special grades of ther-
mally stable epoxy resins; however, there is still a
signicant desire to obtain better thermal expansion
characteristics and delamination performance, espe-
cially in the light of the move to lead-free soldering
which requires higher temperatures.
66
Combinations
of low-brominated novolac and epoxy TBBA have the
advantage of low moisture uptake.
67
Highly thermally stable 2,2

,6,6

-tetrabromo-3,3

,
5,5

-tetramethyl-4,4

-biphenol (Scheme 33) having


the m-brominated phenol moiety was synthesized
and reacted into epoxy resin systems.
68
In electronic
encapsulation and laminate applications, epoxy sys-
tems derived from this brominated biphenol have
exhibited superior hydrolytic and thermal stability as
compared with the conventional o-brominated epoxy
resins. These properties have resulted in an extended
device life for semiconductors and a high glass transi-
tion temperature (T
g
) with excellent blister resistance
for printed circuit boards, while meeting ame retar-
dancy requirements as well.
Low smoke formation and low rate of ame spread
are important factors in structural composite materials.
Bis(hexachlorocyclopentadieno)cyclooctane (Dechlo-
rane Plus) (Scheme 34) in combination with antimony
trioxide provides an efcient ame-retardant level
in composite materials.
69
Alternative synergists, zinc
borate or iron oxide, which partially or completely sub-
stitute for Sb
2
O
3
, allow signicantly reduced smoke
release and make Dechlorane Plus an attractive
replacement for brominated ame retardants, wher-
ever smoke is deemed important.
Scheme 33.
Scheme 34.
The phthalide-containing epoxy resins were syn-
thesized by chain extention with 4,5,6,7-tetrabromo-
phthalein (Scheme 35) and characterized by compar-
ison with TBBA epoxy resins in terms of thermal
properties.
70
Although both resins contain comparable
amounts of halogens, the resulting ame-retardancy
was higher in the phthalide-containing resin. The char
formation upon pyrolysis was also enhanced by the
phthalide functionality.
The ammability, thermomechanical properties and
re response of the diglycidylether of 1,1-dichloro-2,2-
bis(4-hydroxyphenyl)ethylene (DGEBC) (Scheme
36) cured with several hardeners were recently exam-
ined and compared to diglicidylether of bisphenol-
A (DGEBA) systems.
71
The mechanical properties
of the DGEBC and DGEBA systems were equiv-
alent, but the DGEBC systems exhibited superior
ame-resistance and 50 % lower heat-release rate and
heat-release capacity than the corresponding DGEBA
systems. DGEBC cured with methylenedianiline had
an oxygen index of 30, exhibited UL 94V-0/5Vbehav-
ior and easily passed the FAA heat release requirement
FAR 25.853(a-1) as a single-ply glass fabric laminate.
The excellent re-retardant performance of DGEBC
was attributed to its unique charring mechanismrather
than to the effect of chlorine present in the molecule.
Phosphorus-containing additives
Cured epoxy resins have a high concentration of OH
groups,
1,20,21,72
and therefore phosphorus-containing
re-retardants are particularly effective in epoxy resins
because phosphorus-containing products tend to react
with OH groups. Spennger and Utz
73,74
showed that
the amount of phosphorus needed to achieve a V-
0 rating depends strongly on the type of hardener
used, as well as on the presence or absence of bers
or llers. Anhydride hardeners require up to 5 % P,
and dicyandiamide (DICY) usually 3 %. However,
for laminates with 60 % ber content 2 % P can
already be sufcient. As the amount of phosphorus
needed to fulll the ammability requirements is
different for each epoxy application, every system
should be optimized by an iterative test series. Epoxy
resins can be re-retarded by using conventional
additives; however, reactive comonomers often are
more preferred because they allow the maintaining
1908 Polym Int 53:19011929 (2004)
Decomposition, combustion and re retardancy of epoxy resins
Scheme 35.
Scheme 36.
of better physical properties. An extensive review
on phosphorus-containing ame-retardants in epoxy
resins was recently published by Jain et al.
75
The advantage of phosphorus-based printed wiring
boards (PWBs) is that their tracking/arcing perfor-
mance is signicantly improved, although other key
properties may suffer.
65
Water absorption could be
relatively high and likely to cause problems in the man-
ufacture of plated-through holes in the printed circuit
boards. Moisture ingress into the laminate may be
encapsulated by plating down the hole, and can sub-
sequently cause blistering of the plating when heated,
for example, during soldering.
Red phosphorus is a very efcient additive for epoxy
resins; however, it needs to be stabilized in order to
maintain long-term reliability. For example, 4 wt%
of red phosphorus, coated with aluminum trihydrate
(ATH) and encapsulated with phenolic resin, was
used in combination with 25wt% ATH in an adhesive
formulation made from a blend of bisphenol A and
cresol novolac epoxies.
76
A V-0 rating in the UL 94
test was observed with printed circuit board laminates
made with this adhesive. Only 0.5wt%of encapsulated
red phosphorus was required in a cresol novolac resin
lled with 500 parts of silica in order to obtain a V-0
rated resin for packaging of electronic devices.
77
Ammonium polyphosphate (APP) can be used in
some epoxy formulations where long-term hydrolytic
instability can be tolerated.
78
Low smoke generation
of an epoxy containing APP is an advantage.
Combinations with ATH help to improve some
physical properties while maintaining the required
level of ame-retardancy. Although APP shows very
high efciency in epoxy resins (15wt% increases the
OI from 22 to 31 in bisphenol A epoxy cured with an
aliphatic amine), co-addition of the potassium salt of
diphenylsulfonedisulfonic acid helps to boost the OI to
38.
79
Apparently, APP in combination with ATH was
used in early commercial halogen-free printed wiring
boards.
80
Recently, Clariant
81
patented technology on the
use of the aluminum salt of diethylphosphinic acid
(Scheme 37) in combination with ATH or APP
or melamine to ame-retard epoxy adhesives. For
example, use of 10 parts per hundred parts of
rubber (phr) of the aluminum salt and 50 phr of
Scheme 37.
Scheme 38.
ATH resulted in a V-0 rating of the glass-reinforced
laminate. In a patent to Chisso,
82
it was suggested
to partially decompose APP to release 510 % of
the stoichiometric ammonia, and then react the thus-
formed acid groups with melamine. APP coated with
melamine provided a V-0 in bisphenol A type epoxy
at 20wt% loading.
Kodolov et al
83
reported a study on intumescent
composites based on epoxy resins crosslinked with
polyethylenepolyamine and containing APP and
such co-additives as calcium borate, manganese
dioxide and nickel- or chromium-containing tubulenes
(products of dehydration and polycondensation of
phenanthrene containing Ni or Cr) as carbonization
stimulators. In fact, the introduction of metal-
containing tubulenes led to compositions with low
ammability and high char yield, whereas the use of
calcium borate considerably increased the strength of
the intumescent foam being formed.
The ammonium salt of methylphosphonamidic acid
(Scheme 38), encapsulated in polymethyldiethoxy-
siloxane or polyaminopropylethoxysiloxane, was used
as a re-retardant additive in composite materials
containing 60wt% of glass ber reinforcements.
84,85
At 15wt% loading, the composite extinguished in air
within 10s.
The effect of triphenyl phosphate (TPP) and ATH
on the ame-retardancy and thermal stability of
acid anhydride-cured epoxy resin was studied by
thermogravimetry and OI.
86
Unexpectedly, it was
found that ATH, TPP and the mixture of ATH
and TPP all could decrease the char residues of the
cured resin, which is disadvantageous for the ame-
retardant effect. Furthermore, it was shown that this
disadvantageous effect is the most pronounced in the
temperature region of the resins skeletal degradation
(about 360450

C). It was believed that the reactive


alumina (a Lewis acid) which is formed as the ATH
loses its water of crystallization, and the phosphorus
acids formed as the TPP decomposes in the condensed
phase, could catalyze the degradation of the cured
epoxy resin.
A dual thermosetting system consisting of an
epoxy resin, blended with an unsaturated polyester
was ame-retarded either with triphenylphosphine
oxide or various commercial aromatic phosphates
(triphenyl phosphate, isopropylphenyl diphenyl phos-
phate, cresyl diphenyl phosphate, etc.) or diethyl
Polym Int 53:19011929 (2004) 1909
SV Levchik, ED Weil
ethylphosphonate.
87
It was found that in the pres-
ence of a small amount of unreactive phosphorus-
containing additives, the OI increased considerably
but the ame-retardancy was not further increased
by increasing the level of the phosphorus-containing
additive. In fact, a loading of 5 % of the additives
gave almost the same OI as 15 %. The presence of
phosphorus in the formulation contributed to the
formation of an intumescent char. Only 7.5 wt%
TPP was necessary to provide a V-0 rating in a
novolac epoxy/bismaleimide blend cured by 4,4

-
diaminodiphenylmethane.
88
A comparison of DGEBA/DDS resins containing
phosphate compounds as additives, namely, trimethyl
phosphate (TMP), triethyl phosphate (TEP), tri-
butyl phosphate (TBP) and triphenyl phosphate
(TPP), with resins including chemically bonded
organophosphate groups, namely, DGEBA pre-
reacted with diphenyl or dialkylester phosphates, was
performed by Derout et al.
89
They showed that the
re-retardant behavior of crosslinked DGEBA/DDS
with incorporated dialkyl phosphate groups onto the
epoxy resin backbone was always greater than those
of the resins containing the corresponding trialkyl
phosphate additives. Moreover, phenyl phosphate
derivatives lead to a good increase of re retardancy
(OI > 28): the comparison with the OI value
of a DGEBA/DDS specimen containing triphenyl
phosphate as additive proves that the best re behavior
was achieved when phenyl phosphate molecules are
chemically bonded onto the resin backbone.
Triphenyl phosphine oxide at 10wt% was found
to provide a V-0 UL 94 rating in the printed circuit
board laminates manufactured with high-performance
epoxy resins such as the blends of cresol novolac,
phenolic alkyl novolac, triazine-modied novolac and
xylene-modied novolac epoxy resins.
90
Liu et al
91
showed that the re-retarding proper-
ties (OI and UL-94) of a cured epoxy were strongly
improved by using the commercial additive resor-
cinol bis(diphenyl phosphate) (RDP). Moreover, a
combination of the phosphate with phenolphthalein,
used as a chain extender, gave substantially enhanced
ame-retardant results. A related system where the
phenolphthalein structure was built into the phosphate
additive (Scheme 39), rather than into the polymer
backbone, gave inferior results. Thermogravimetric
analysis (TGA) and infrared spectroscopy of the solid
residue suggested that the phenolphthalein group is
involved in a crosslinking reaction, which may lead
to char enhancement. The RDP depressed somewhat
both the glass transition temperature (T
g
) and ther-
mal stability, but the presence of the phenolphthalein
structure partially overcame the T
g
depression.
A series of hindered bisdiphosphates (Schemes 40
42) in combination with 80 % of amorphous silica was
tested in an epoxy formulation for the encapsulation
of electronic devices.
92
A V-0 rating in the UL-94
test was observed at 1.5 % phosphorus content in the
epoxy adhesive.
Scheme 39.
Scheme 40.
Scheme 41.
Scheme 42.
Scheme 43.
Scheme 44.
Unexpectedly, antimony trioxide, a primary syn-
ergist for halogen-containing ame-retardants but
rarely effective with phosphorus, was found to be
synergistic with bisphenol A tetraxylenyl phosphate
(Scheme 43).
93
For example, 16wt% of the phos-
phate and 11wt% of Sb
2
O
3
provided a V-0 rating in
the blend of novolac epoxy and bis(cyclohexyl)propane
epoxy (Scheme 44) lled with 10wt% silica.
1910 Polym Int 53:19011929 (2004)
Decomposition, combustion and re retardancy of epoxy resins
Scheme 45.
Scheme 46.
Scheme 47.
Epoxy resins containing dimethyl methylphospho-
nate (DMMP), which aids in the processing of the
epoxy by reducing the viscosity of the resin mixture,
were cured with amines.
94
DMMP also acts as an
antiplasticizer in the cured epoxy by increasing the
modulus and yield strength. The ammability was
investigated by using a microcalorimeter, which mea-
sures the amount of oxygen consumed during pyrolysis
and provides a quantitative evaluation of the heat
released upon combustion. The total heat released
was unchanged by the addition of the additive, hence
indicating that the mechanism for degradation was
not changed with the addition of DMMP. The lack
of additional char formation is also an indication of
this. However, the heat-release capacity is reduced in
a manner similar to the decrease of degradation rate
seen by TGA, from a peak of 1063 to 365J g
1
K
1
at
15 phr DMMP.
A phosphonate salt, made by the prolonged heating
of DMMP with urea, was added at an 11wt% level
to DGEBA, and the resin, then cured with phenol-
formaldehyde novolac, showed a V-0 rating.
95
The
aluminum salt (Scheme 45), made by reacting methyl
methylphosphonic resin with ATH, provided a V-0
rating at 24wt% loading.
96
Salts made by heating
DMMP with salts of Mg, Ca, Zn, Ba or Sb were
less efcient and showed only a V-2 rating at 25wt%
loading.
97
Epoxy-based laminates containing 14wt% of the
commercially available cyclic phosphonate (Scheme
46) showed a V-0 rating in the UL 94 test.
98
Similar
cyclic phosphonates (Scheme 47), where R is butyl,
hexyl, octyl or 1,2-ethandiy/hexyl (oligomeric) groups,
showed a V-0 rating at 25wt% loading.
99
The inuence of different amounts of poly(propoxy-
phosphazene) (Scheme 48) on the curing kinetics,
physical properties and ame-retardancy of bisphenol
A type-epoxies cured with diethylenetriamine was
investigated by Denq et al.
100
The results revealed that
the polyphosphazene was partially miscible with the
epoxy and could be a catalyst or a diluent, depending
Scheme 48.
on the content. The tensile strength and the modulus
of the blends decreased with increasing amounts of the
polyphosphazene, although the elongation increased
with increasing ame-retardant (FR).
Mirkamilov and Mukhamedgaliev
101
phosphory-
lated a gossipol resin, which was a waste product
of the fat and butter industry containing signicant
amounts of phenolic functionalities in its structure.
Phosphorylation was carried out by using either phenyl
phosphates or phosphoramidates. It was found that the
time of ignition of the specimens and the OI values
signicantly increased upon addition of the phospho-
rylated resin to the epoxy. The increase in the content
of H
2
O and CO in the gas phase of the combustion
products was accompanied by a signicant reduc-
tion in the concentration of hydrocarbons. This was
attributed to the fact that the burning gases undergo
oxidative decomposition inside the pre-heating zone
of the diffusion ame before they can reach the ame
front. However, the modied specimens also showed
an increase in soot formation.
The use of tris(dichloropropyl) phosphate (TCPP)
as a ame retardant in epoxy resins cured by
methyltetrahydrophthalic anhydride was studied by
Li et al.
102
The effect of the phosphorus content
of the TCPP on the OI of the epoxy resins was
compared with the results obtained when using
triphenyl phosphate and trimethylphenyl phosphate.
Although TCPP is an organophosphorus ame-
retardant containing chlorine, it was not more effective
when compared to other ame retardants containing
only phosphorus. It appeared quite possible that
TCPP mainly works through phosphorus alone,
despite the presence of halogen.
Maltseva et al
103
measured the temperature on the
back side of the sample heated at 10Wm
2
with a
laser. The plots showed complicated shapes: a sharp
increase of the temperature at the initial moment
corresponded to heating of the composition before a
char layer was formed, then an inection of the curve
related to the formation and growth of the char was
observed, and nally, a slow temperature increase,
up to a stationary state, was indicative that the char
formation process was nished. It was found that non-
ame-retardant epoxy resin produces char slowly, and
that the protective properties of the char are low,
whereas epoxy resins with phosphorus-containing FRs
typically showed higher rates of char formation and
better thermal insulative properties.
Phosphorus-containing epoxy monomers
There is a relatively large number of publications about
reactive phosphorus-containing epoxy monomers.
Polym Int 53:19011929 (2004) 1911
SV Levchik, ED Weil
Scheme 49.
Scheme 50.
Series of diglycidyl phosphates, diglycidyl phospho-
nates and glycidyl phosphinates have been prepared
in the laboratories of Siemens AG.
104
It was found
that such phosphonates and phosphinates are more
efcient at the same level of phosphorus content than
those phosphates, either incorporated into the resin
network, or added (tricresyl phosphate).
There are a large number of publications discussing
the use of 9,10-dihydro-9-oxa-10-phosphaphenan-
threne 10-oxide (DOPO) (Scheme 49). This can be
either pre-reacted with epoxy resin or used as a reactive
additive during curing. The glass-ber-reinforced
laminates prepared with DOPO could be classied
V-1 with 1.6 % phosphorus and V-0 with 2.1 %
phosphorus.
73,74,105,106
With proper formulation, the
stress at failure of these phosphorus-containing ame-
retarded laminates, as well as their glass transition
temperatures, can be similar to non-ame-retardant
epoxies. The modulus and the strain at failure are
comparable with standard laminates. Investigations of
the chemical resistance of DOPO-containing epoxy
systems showed no large difference when compared
to the control, except for resistance to ammonia. The
weight gain when immersed in aqueous ammonia was
signicant for phosphorus contents above 23 %.
Phosphorus-containing epoxy resins (Scheme 50)
(13 % phosphorus content) were synthesized by
reacting DOPO with DGEBA.
107,108
Higher OI values
were obtained with higher phosphorus contents. For
resin cured with 4,4

-aminodiphenyl sulfone (DDS),


the OI increased from 22 to 28 when the phosphorus
content increased from 0 to 1.6 %. For the phenolic
novolac (PN) curing system, the OI increased from 21
to 27 when the phosphorus content increased from 0
to 2.2 %. A V-0 rating was achieved with both curing
agents.
Similar monomers (Scheme 51) were prepared by
reacting DOPO with cresol formaldehyde epoxy
novolac resin (average functionality, 12).
109,110
The
onset degradation temperatures, as measured by TGA,
decreased with phosphorus content; however, com-
paring with other phosphorus-containing polymers,
the decomposition temperatures of these phosphorus-
containing epoxies, were relatively high. Char yields
increased with phosphorus content. Although incor-
porating DOPO into epoxy resins reduced their T
g
Scheme 51.
Scheme 52.
values and crosslink densities, their T
g
values were still
higher than for other phosphorus-containing advanced
epoxy resins used in FR-4 circuit board applications.
The OI values increased from 23 to 27 when the
phosphorus content was increased from 0 to 1.7 %,
and increased to 33 when the phosphorus content was
3.6 %. The OI for dicyandiamide (DICY)-cured resin
was 34, which was higher than for DDS- or PN-cured
resins. This was attributed to the high nitrogen content
in DICY (N = 67 %). A similar trend was observed
for the UL-94 test.
A series of advanced epoxy resins with various
epoxy equivalent weights were synthesized from a
reactive phosphorus-containing diphenol (Scheme 52)
and diglycidyl ether (Scheme 53) and then cured with
DDS, PN or DICY.
111,112
It was shown that the
decomposition temperatures of the formulations with
different phosphorus contents are almost the same as
for non-ame-retardant epoxies. Less than 1 % char
yield was found in the non-ame-retardant system
at 700

C under air; however, 1018 % char yields


were found in the phosphorus-containing epoxies.
The OI increased from 22 to 28 when the phosphorus
content increased from0 to 1.4 %; however, it reached
a plateau when the phosphorus contents exceeded
2.1 %. In the epoxy formulation for encapsulation
of electronic devices, a V-0 rating could be achieved
with a phosphorus content of 1.0 %, comparable to a
bromine content of 7.2 %.
In another study, the above-mentioned diphenol
(Scheme 52) was used as a reactive ame retardant in
o-cresol formaldehyde novolac epoxy resin.
106,113,114
Because of the rigid, cyclic, side-chain structure of
DOPO, the resultant phosphorus-containing epoxy
resin exhibited a higher glass transition temperature,
better ame retardancy, higher modulus and greater
thermal stability than the regular bromine-containing
tetrabromobisphenol A epoxy resin. A UL 94V-0
1912 Polym Int 53:19011929 (2004)
Decomposition, combustion and re retardancy of epoxy resins
Scheme 53.
Scheme 54.
rating was achieved with a phosphorus content as low
as 1.1 %, comparable to a bromine content of 613 %.
The same diphenol (Scheme 52) and a similar
naphthalene-type material (Scheme 54) have been
copolymerized with novolac-type epoxy resins.
115
A
V-0 UL 94 rating was achieved in printed circuit
board laminates made with this epoxy at 1.52.0 %
phosphorus in the resin. A V-0 rating was observed
in cresol novolac epoxy formulations if 25 % of
the phenol-formaldehyde novolac used to cure the
epoxy was substituted with the diphenol, dinaph-
thol (Schemes 52 and 54) or their related methyl-
substituted compounds (Schemes 55 and 56).
116
Other ame-retardant epoxy monomers were
prepared by reacting the diphenol (Scheme 52)
with DGEBA (Scheme 57) or high-performance
naphthalene-based epoxies (Scheme 58) and com-
pared with TBBA (Scheme 32) advanced epoxy
resins.
117
A very high glass transition temperature
Scheme 55.
Scheme 56.
(T
g
= 235

C) was obtained when the tetrafunctional


naphthalene-containing epoxy resin was used. The OI
values increased with the increase of bromine con-
tent (from 32 to 39 when the bromine content was
increased from 13.4 to 22.7). Phosphorus was found
to be more effective than bromine in attaining the
ame-retardant property (2 % phosphorus seemed to
be as effective as 20 % bromine). A V-0 rating could
be achieved with 1.4 wt% phosphorus or 13.4wt%
bromine for highly crosslinked resins.
Two phosphorus-containing diacids were synthe-
sized from DOPO (Scheme 49) and either maleic acid
(Scheme 59) or itaconic acid (Scheme 60) and then
reacted with the diglycidyl ether of bisphenol Ato form
two series of advanced epoxy resins.
118
After curing
with DDS, the thermal properties of the cured epoxy
resins were studied by using dynamic mechanical anal-
ysis (DMA) and TGA, while the ame-retardancies
of the cured epoxy resins were evaluated by using
Scheme 57.
Scheme 58.
Polym Int 53:19011929 (2004) 1913
SV Levchik, ED Weil
Scheme 59.
Scheme 60.
Scheme 61.
Scheme 62.
Scheme 63.
the UL-94 test. The degradation temperatures were
found to decrease with phosphorus content, while
the char yield increased with phosphorus content.
The itaconic-acid-based formulations required slightly
higher phosphorus content in order to achieve a rating
in the UL-94 test. For both types of ame-retardant
epoxies, incorporating 1.7 % P provided a V-1 rating.
Dicarboxylic acids, containing either aliphatic
(Scheme 61) or aromatic (Scheme 62) phosphine
oxides, or DOPO-itaconic acid structures (Scheme
60), were copolymerized with DGEBA at 22 and
30wt%, respectively.
119
When these copolymers were
cured with DICY, epoxy resins with a V-0 UL-94
rating were obtained.
DOPO (Scheme 49) was reacted with epichloro-
hydrin (Scheme 63) and then polymerized to give
a linear polyether-type prepolymer (Scheme 64).
120
This prepolymer could not be crosslinked when using
amines as hardeners; however, cationic initiators or
acid anhydride hardeners were suitable for obtaining
epoxy resins with high phosphorus contents and good
ame-retardant properties.
DOPO (Scheme 49) was reacted with the car-
bonyl group in 4,4

-dihydroxybenzophenone to give
a phosphorus-containing bisphenol (Scheme 65),
Scheme 64.
Scheme 65.
Scheme 66.
which was then reacted with epichlorohydrin or
DGEBA to obtain phosphorus-containing epoxy
monomers.
121123
Differential scanning calorimetry
(DSC) and thermogravimetric analysis revealed that
these cured epoxy resins had high glass transition
temperatures and high thermal stabilities. The char
yields were found to increase when the phosphorus
content increased, while increasing the phosphorus
content resulted in increasing the OI values. The
DICY-cured epoxy resins showed higher OI values
when compared to other diamines. This was attributed
to a possible phosphorusnitrogen synergism. When
the DOPO-containing epoxy resins were cured with
a DOPO-containing diamine (Scheme 66), extremely
high OI values, between 37 and 50, were found, result-
ing from the high phosphorus contents of the epoxy
resins.
Glycidyloxy diphenylphosphine oxide (Scheme 67),
a side-chain phosphorus-containing monoepoxide,
was mixed with DGEBA and cured with diethylene-
tetramine.
124
Because of the low crosslink density
and gel fraction of the system with the monoe-
poxide, this resulted in poor protection by intu-
mescent char during thermal decomposition, and
therefore the OI of the epoxy decreased with an
increase of the monoepoxide, although the phos-
phorus content increased. In contrast, the OI of
the polymer increased with the phosphorus con-
tent of the samples when DGEBA was copoly-
merized with bis-(3-glycidyloxy)phenylphosphonine
oxide (Scheme 12).
125,126
These systems showed rel-
atively higher char yield and OI values at 2931,
1914 Polym Int 53:19011929 (2004)
Decomposition, combustion and re retardancy of epoxy resins
Scheme 67.
Scheme 68.
while the epoxy without phosphorus exhibited OI
values of 1922. The same resin was copolymer-
ized with bis(4-aminophenyl) phenylphosphate and
showed a much higher OI of 3439.
127
Bis-(3-
glycidyloxy)phenylphosphonine oxide has also been
used to prepare an interpenetrating network resin
in combination with triallylisocyanurate, which was
cured with a peroxide.
128
The resin containing 2 %
phosphorus extinguished immediately after removal
of the ame.
Aliphatic diglycidyl phosphonates (Scheme 68),
where R is CH
3
(Scheme 68a),
129,130
C
3
H
7
(Scheme
68b)
130,131
or C
6
H
6
(Scheme 68c)
130
were copolymer-
ized with DGEBA and novolac epoxy resins and then
cured with bis(4-aminophenyl) ethylphosphine oxide.
These adhesives provided V-0 ratings for laminate
printed wiring boards
131
or S4/SR2 rating in the DIN
551-2 test.
130
Phenylphosphonic dichloride was rst reacted with
two moles of a bisphenol and then the dihydroxy
compound was converted to an epoxy monomer
by reaction with epichlorohydrin (Scheme 69).
132
The OI values of these phosphorus-based epoxies
(Schemes 69a69d) ranged from 33.3 to 40.3. The
maximum OI value was obtained for the epoxide
with the highest P content (Scheme 69b), while
the minimum OI value was exhibited by the
epoxide containing the sulfur moiety (Scheme 69d).
Chemically bound sulfur was found to show no
contribution to the ame-retardancy of the polymer.
One of these epoxy resins (Scheme 69a) was further
co-reacted with an s-triazine derivative, thus produc-
ing a phosphorusnitrogen-containing monomer.
132
This epoxy was more efcient than the purely
phosphorus-containing resin, hence indicating an
N/P synergistic effect. This synergistic effect showed
a maximum at a P/N ratio of 1.8. Similarly,
epoxy monomers were prepared with phenolph-
thalein phosphorus-containing groups (Scheme 70)
and s-triazine nitrogen-containing groups. This epoxy
showed a maximum synergistic performance at a P/N
ratio of 1.2. It was speculated that steric hindrance
of the phenolphthalein group slows down the forma-
tion of the PN network structure responsible for the
synergism.
An epoxy resin containing a cyclic phosphine
oxide group in the main chain (Scheme 71)
133
(a)
(b)
(c)
(d)
Scheme 69.
Scheme 70.
Scheme 71.
Scheme 72.
and a tetrafunctional epoxy containing phosphorus
oxide and nitrogen groups in the main chain
(Scheme 72)
134
were synthesized and cured with
bis(3-aminophenyl)methylphosphine oxide (BAMP),
4,4

-diaminodiphenylmethane (DDM) or 4,4

-di-
aminodiphenyl sulfone (DDS), and compared with
commercial resins. In any combination, either cured
with phosphorus-containing BAMP, DDM or DDS,
the phosphorus-containing epoxy resins showed
higher char yields than bisphenol A or novolac-
type commercial resins. This was extrapolated to the
high ame-retardant properties of the phosphorus-
containing resins.
An oligomer prepared by reacting bisphenol A with
phenylphosphonic dichloride (Scheme 73) was used
for chain extension of DGEBA and compared with
commercial TBBA chain-extended epoxies.
135
Epoxy
resins with polyphosphonates, containing 1.51.8 %
Polym Int 53:19011929 (2004) 1915
SV Levchik, ED Weil
n
Scheme 73.
CH
3
C
2
H
5
C
4
H
9
(a)
(b)
(c)
(d)
Scheme 74.
Scheme 75.
P, exhibited a lower ammability than the analogues
without P, although the ammability was still higher
than that of the commercial brominated resin
containing 18.5 % Br.
By reacting DGEBA with dialkyl (or aryl) phos-
phates, it was possible to chemically modify the epoxy
resin (Scheme 74) and then cure it with DDS to
obtain a resin with good ame retardancy and ther-
mal stability.
89
Chemical modication of DGEBA by
dialkyl (or aryl) phosphates could be carried out in situ
during the curing of epoxy resins without any change
in re behavior. The re-retardancy of epoxy resins
did not signicantly improve by chemical introduction
of dialkyl phosphate grafts (Schemes 74a74c) at a
10 % level. On the other hand, the phenyl phosphate
derivatives (Scheme 74d) lead to a good decrease in
the ammability (OI > 28).
Novolac epoxy resins were phosphonylated with
phosphonic anhydrates (Schemes 7577).
136,137
Pri-
nted circuit board laminates prepared with epoxy
containing 2.7 % phosphorus exhibited a UL 94V-
0 rating. Methylethyl- (Scheme 78) or dimethylpy-
rophosphinates (Scheme 79) were used for the phos-
phorylation of phenolic novolac epoxy resin, which
ensured its ame retardancy.
136
Aliphatic pyrophos-
phates (Schemes 8082) or pyrochlorophosphates
(Schemes 83 and 84) were made by reacting aliphatic
phosphate or chloroalkyl phosphate with P
4
O
10
.
138
Epoxy resins pre-reacted with 12wt% pyrophosphates
were cured with a polyamine in order to prepare coat-
ings. These coatings passed the DIN 4102 ammabil-
ity test.
A phosphorus-containing epoxy resin, bis(3-tert-
butyl-4-glycidyloxyphenyl-2,4-di-tert-butylphenyl)
resorcinol diphosphate (Scheme 85), was synthesized
Scheme 76.
Scheme 77.
Scheme 78.
Scheme 79.
Scheme 80.
Scheme 81.
Scheme 82.
Scheme 83.
Scheme 84.
1916 Polym Int 53:19011929 (2004)
Decomposition, combustion and re retardancy of epoxy resins
Scheme 85.
Scheme 86.
and subsequently cured with aromatic amines.
139
The
phosphorus-free epoxy polymers exhibited OI values
of 1823, whereas the phosphorus-containing epoxy
(Scheme 85) cured with commercial amines exhib-
ited values of 2831. The larger OI values could be
obtained for the phosphorus-containing polymers with
high aromatic contents. This was attributed to the char
contribution of the phenyl groups.
A multifunctional epoxy resin was prepared
by the reaction of epichlorohydrin with (4-
diethoxyphosphoryloxyphenoxy)(4-hydroxyphenoxy)
cyclotriphosphazene (Scheme 86).
140
The epoxy resin
was further cured with diamine curing agents, DDM,
DDS, DICY, and 3,4

-oxydianiline (ODA). Com-


pared to DGEBA, epoxy polymers (Scheme 86)
showed lower weight-loss temperatures, higher char
yields, and higher OI values, hence indicating that
the epoxy resin thus prepared could be useful as a
ame-retardant.
Phosphorus-containing curing agents
Similarly to epoxy monomers, phosphorus-containing
groups can be incorporated into the curing
agent. 9,10-dihydro-9-oxa-10-phosphaphenanthrene
10-oxide (DOPO) (Scheme 49) is commercially avail-
able in Japan and Europe and is recommended
as part of a curing system for halogen-free ame-
retardant epoxies.
141
Because of the latency of the
curing agent, combination resins offer the easy han-
dling of one-component systems. Because the ame-
retardant content can be freely varied, the systems
could be adapted to specic ame-retardancy require-
ments. Combination resins containing DOPO were
found to meet the UL 94V-0 rating with phospho-
rus contents below 3 %. This meets the require-
ments for electrical laminates. However, a study
using a cone calorimeter showed that a test lam-
inate containing this additive still does not meet
the ame-retardancy requirements for aerospace
applications.
Methylsuccinic anhydride was reacted with DOPO
and the product (Scheme 87) was used as a reactive-
type ame-retardant with a pendent phosphorus group
Scheme 87.
Scheme 88.
to cure DGEBA and epoxy-novolac.
142
Epoxy resins
cured with this phosphinate anhydride exhibited
higher OI values and char yields when compared
with those cured by commercial anhydrides, such as
phthalic anhydride and hexahydrophthalic anhydride.
Surprisingly, the phosphinate anhydride showed a
better effect on promoting re-resistance in DGEBA-
based resins than in epoxy novolac-based resins at the
same phosphorus content.
A product of the interaction of DOPO and p-
hydroxyphenylmaleimide (Scheme 88) was reacted
with epoxy and cured by phenol-formaldehyde
novolac.
143,144
The resin thus obtained was useful
for encapsulating electronic devices and gave V-
0 ratings in the presence of a high content of
silica.
A phosphorus-containing novolac was prepared
from DOPO and 4-hydroxybenzaldehyde via an
addition reaction (Scheme 89).
145
Since this product
contained multi-phenol groups in the molecular chain,
it was used as a polyfunctional curing agent. High glass
transition temperatures (above 160

C) and very high


thermal stabilities (above 300

C) were observed for


o-cresol novolac epoxy cured by using this materials
(Scheme 89). The OI values were proportional to the
phosphorus contents of the epoxy resins. A minimum
phosphorus content of 2wt% was found to enable the
cured epoxy resins to exhibit an OI value of 26 and to
pass the UL 94 test at V-0.
Polym Int 53:19011929 (2004) 1917
SV Levchik, ED Weil
Scheme 89.
Scheme 90.
In another study, this phosphorus-containing
novolac (Scheme 89) was compared to the prod-
uct of partial phosphorylation of phenol formalde-
hyde novolac with diphenyl phosphorochloridate
(Scheme 90).
146
Owing to the rigid structure of
the former (Scheme 89), the resultant phosphorus-
containing epoxy resin exhibited a higher glass
transition temperature, and thermal stability than
the phosphorylated phenol formaldehyde novolac
(Scheme 90)-containing epoxy or TBBA-containing
epoxy resin. By comparing the efciencies of these
two resins (Schemes 89 and 90), it was found that the
epoxy resins cured with the phosphorus-containing
novolac (Scheme 89) demonstrated higher OI values
and shorter burning times in the UL 94 test than
phosphorylated novolac (Scheme 90) (1.35 % P in the
phosphorus-containing novalac (Scheme 89) is better
than 1.71 % P in the partially phosphorylated material
(Scheme 90)). These results were in good agreement
with the char yields.
A phosphorus-containing aralkyl novolac was pre-
pared from the reaction of DOPO, rst with tereph-
thaldicarboxaldehyde and subsequently with phenol
(Scheme 91).
147
This product blended with phe-
nol formaldehyde novolac was used as a curing
agent for o-cresol formaldehyde novolac epoxy. The
epoxy resins exhibited high glass transition tempera-
tures (15977

C), good thermal stabilities (>320

C)
and low thermal degradation rates. High char yields
and high oxygen index values (2632.5) were
observed, hence indicating that the resins had good
ame-retardance. Using a melamine-modied phenol
formaldehyde novolac to replace phenol formaldehyde
novolac in the curing composition further enhanced
the glass transition temperatures (160186

C) and
OI values (2833.5) of the cured epoxy resins.
DGEBA was cured with a phosphorus-containing
diamine derived fromDOPOand 3-nitrobenzoyl chlo-
ride, followed by hydrogenation (Scheme 92).
148
This
Scheme 91.
Scheme 92.
Scheme 93.
Scheme 94.
Scheme 95.
resin was compared with the phosphorus-free epox-
ies cured with DDS or DDM. The char yield of
the phosphorylated epoxy at 700

C under nitro-
gen was 32 %, whereas the yields of DGEBA/DDS
and DGEBA/DDM were 15 % and 13 %, respec-
tively. This result was further conrmed by OI
measurements. Compared with the OI values of
DGEBA/DDS (22) and DGEBA/DDM (21), the OI
of the phosphorus-containing epoxy was 30. Corre-
lations between the high char yields and OI values
implied that the ame-retardancy of the epoxy resins
was elevated by using the phosphorus-containing
diamine as a curing agent.
The synthesis of the diamines containing two DOPO
groups (Schemes 9395) has been described by Chiu
et al.
123
1918 Polym Int 53:19011929 (2004)
Decomposition, combustion and re retardancy of epoxy resins
Scheme 96.
Scheme 97.
Phosphine oxide structures are often used to impart
ame-retardancy to curing agents, because phosphine
oxides are thermally and hydrolytically very stable. A
cyclic phosphine oxide (Scheme 96) was used in vari-
ous epoxy resins, including the naphthalene type and
branched novolac epoxies.
115
Most of these compo-
sitions showed a V-0 rating in printed circuit board
laminates at 1.52.0 % phosphorus in the epoxy.
A few studies have been reported in the
literature concerning curing of epoxy resins with
bis(aminophenyl)methylphosphine oxide (Scheme
97), as well as the thermal and combustion
performance of the resins containing this
oxide.
56,87,133,149153
Von Gentzkow et al
129
have used
the analogous bis(aminophenyl)ethylphosphine oxide
(Scheme 98) to cure a phosphorus-containing epoxy
resin.
Varma and Gupta
149
found that glass-fabric-
reinforced laminates based on DGEBA and the
bis(aminophenyl)methylphosphine oxide (Scheme 97)
exhibited a higher limiting oxygen index, as well
as higher shear and exural strength, than those
based on the DGEBA/DDS system. Thermal ageing
at 185

C for 100h did not affect the mechanical


properties of the phosphorus-containing laminates;
however, a signicant decrease was observed in the
interlaminar shear strength by boiling in water for 100
and 200h. Epoxy resins cured with this oxide (Scheme
97) exhibited higher char yields when compared with
those cured by DDS.
153
The phosphorus-containing
epoxy resins showed self-extinguishing characteristics
in comparison to the DDS-cured sample which
continued to burn after removal of the ignition source.
The cone calorimetry test showed that the heat-
release rate decreased as the phosphorus concentration
increased.
Shau and Wang
133
compared this methylphos-
phine oxide (Scheme 97) with 4,4

-diaminodiphenyl-
methane (DDM), and DDS in bisphenol A
and novolac epoxy resins, as well as with a
phosphorus-containing epoxy with a cyclic phos-
phine oxide group (see previous section). Simi-
lar comparisons were made with diimidediepoxide
polymers.
56
Epoxy resins cured with the methylphos-
phine oxide (Scheme 97) showed higher OI val-
ues than non-phosphorus-containing resins; however,
Scheme 98.
Scheme 99.
Scheme 100.
the decomposition temperatures of the phosphorus-
containing polymers were signicantly lower. The
authors speculated that phosphoric acid species were
formed upon thermal decomposition, which catalyze
dehydration of the polymers at lower temperatures.
Levchik et al
150
showed that the re-retardant effec-
tiveness of the methylphosphine oxide (Scheme 97)
goes through a maximum with increasing phospho-
rus concentration in a tetraglycidylmethylenedipheny-
lamine (TGMDA)/DGEBA epoxy blend. This was
attributed to the competition between the char-
forming re-retardant action and promotion of the
evolution of combustible gases because of the catal-
ysis of degradation. The second action seems to
prevail at larger concentrations of the oxide, corre-
sponding to 22.5 % P. This is also the range of
concentration where the mechanism of action of the
oxide switches from a condensed-phase (charring) to
a gas-phase action, probably due to volatilization of
ame-inhibiting phosphorus-containing degradation
products.
In line with the other phosphorus-containing epoxy
resins discussed above, it was found that com-
pletely aromatic bis(aminophenyl)phenylphosphine
oxide (Scheme 99) tends to produce more char than
bis(aminophenyl)methylphosphine oxide (Scheme
97).
153
It also gives higher char yields compared to
bis(aminophenyl)phenylphosphonate (Scheme
100).
154
Hsiue et al
155
studied the ame retardant
properties of this phosphonate in combination with
phosphorus- or silicon-containing epoxies and found
a phosphorussilicon synergistic effect.
In another study, the diamine bis(4-aminophenyl)
phenylphosphate (Scheme 101) was studied in combi-
nation with diglycidyl phenylphosphate
39
(see previous
section). High char yields (3252 %) as well as high
oxygen index values (3449) for these phosphorylated
resins were found. TGA studies showed that the
Polym Int 53:19011929 (2004) 1919
SV Levchik, ED Weil
Scheme 101.
Scheme 102.
decomposition of the phosphate groups occurred
independently at relatively low temperatures. The
decomposition of the phosphate groups resulted in
a phosphorus-rich residue at the initial stage and
slowed down further decomposition of the resins.
This not only decreased the decomposition rates in
the high-temperature regions but also resulted in high
char yields.
Various amounts of bis(3-hydroxyphenyl) phenyl
phosphate (Scheme 102) were added to phenol
novolac as a curing agent for DGEBA.
156
Raising
the phosphorus content of the resin system from 0
to 2.4 % increased the char yield from 18 to 35 %.
The UL 94 test showed that the ame-retardancy
of the cured epoxy resins increased with phosphorus
or bromine content (TBBA (Scheme 32)); however,
phosphorus was found to be more effective than
bromine as a ame-retardant (1 % phosphorus is
better than 6 % bromine). In order to achieve a V-0
rating in o-cresol formaldehyde novolac epoxy resins,
2.2 wt% phosphorus or 12.9wt% Br was required.
114
Phosphorus-containing samples generated much less
smoke than those containing bromine.
Di(hydroxy-o-tert-butyl)phenyl phenyl phosphate
(Scheme 103) was used at 30wt%loading in DGEBA,
which helped to increase the OI from 18.5 to
26.5.
157
The use of tris(3-aminophenyl)phosphine oxide
(Scheme 104) as a curing agent for phosphorylated
epoxyimide polymer and commercial bisphenol A
and novolac epoxy resins has been reported.
55
It was
Scheme 103.
Scheme 104.
Scheme 105.
found that the phosphine oxide increases the char
yield for both commercial epoxy resins, whereas the
phosphorylated epoxyimide polymer gives a further
boost to the char yield.
Tris(hydroxytolylphosphine oxide) (Scheme 105)
was prepared by the isomerization of tricresyl phos-
phate using an extremely strong base.
158
Laminates
prepared with novolac epoxy resin cured with 57wt%
of the phosphine oxide showed a V-0 rating.
A poly(arylene ether surfone) phosphine oxide
(Scheme 106), with controlled molecular weights
and amine end groups, was synthesized, and used
as a modier for DGEBA-based epoxy resins.
159
The closely related poly(arylene ether sulfone)
(Scheme 107) was utilized for comparison purposes.
The epoxy samples modied with 20wt% of the
phosphorus-containing ether showed a burning time
of 2030s after removal from the ame, whereas the
epoxy control samples burned up completely.
Scheme 106.
Scheme 107.
1920 Polym Int 53:19011929 (2004)
Decomposition, combustion and re retardancy of epoxy resins
tC
4
H
9
tC
4
H
9
tC
4
H
9
tC
4
H
9
Scheme 108.
Scheme 109.
Scheme 110.
Two hydroxy-terminated oligomeric phosphates
(Schemes 108 and 109) were used alone or in
combination with commercial phosphine oxide
(Scheme 110) or with the aluminum salt of methyl
methylphosphonic acid (Scheme 45).
160
Novolac-type
epoxy resins, heavily loaded with silica (82wt%),
required only 1wt% of the oligomeric phosphate and
0.5 wt% of the co-additive in order to ensure a V-0
rating.
A series of phenylphosphonic acid amides (Schemes
111113) and an amide of phenyl phosphoric acid
(Scheme 114) were evaluated as curing agents for
epoxy resins.
161
The char yield at 800

C, measured by
thermogravimetry, was used as an indicator for ame
retardancy of these curing agents. The amide shown in
Scheme 114 has a lower phosphorus content (4.81 %)
than that shown in Scheme 112 (4.93 %); however,
the char yield of the epoxy cured with the former
(Scheme 114) was signicantly higher than that of the
polymer cured with the latter amide (Scheme 112)
(30.3 % vs 15.9 %). Moreover, the temperature of
the beginning of decomposition epoxies cured with
the amide of phenyl phosphoric acid (Scheme 114)
was also higher than that for other amide-cured
(Scheme 111) epoxies. When companing the epoxies
cured with the amide shown in Scheme 113 to
those cured with the amide shown in Scheme 112,
it was observed that the former has a higher onset
temperature for thermal decomposition but a lower
char yield, which indicates that the presence of the
additional aromatic group in this amide (Scheme 113)
did not play a signicant role in improving the ame
retardancy.
Scheme 111.
Scheme 112.
Scheme 113.
Scheme 114.
Scheme 115.
Scheme 116.
Polym Int 53:19011929 (2004) 1921
SV Levchik, ED Weil
Scheme 117.
Scheme 118.
Scheme 119.
Scheme 120.
Scheme 121.
Bertram and Davis
162
prepared epoxy formula-
tions cured by phenol-formaldehyde novolac resin
and aromatic or aliphatic phosphoric acid amides
(Schemes 115 and 116). These epoxy resins showed
improved ame-retardancy.
Two series of novel phosphorus-containing poly-
alkylene amines, with or without aromatic groups,
were synthesized via reacting phosphoryl chloride
derivatives with commercially available polyether-
amines, ethylenediamine and N-phenyl-1,4-phenyl-
enediamine (Schemes 117122).
139,163
The phos-
phorus-free epoxy polymers were found to have OI
values of 1821, whereas the OI values of the
phosphorus-containing polymers leveled out at 2231.
A relatively high OI value was obtained for the
epoxy cured with the amine shown in Scheme 119,
which indicates the role of aromatics in charring.
The high phosphorus content also resulted in higher
OI values. For example, the polymers shown in
Scheme 118 (P content = 3.46 %) and Scheme 121
(P content = 4.07 %) displayed relatively high OI
values of 27 and 31, respectively.
Liu et al
164
designed and synthesized a highly
stable phosphine oxide containing a secondary
Scheme 122.
Scheme 123.
Scheme 124.
diamine, bis(4-cyclohexylaminophenyl) phenyl phos-
phine oxide (Scheme 123). The diamine, along with
DDS, was then used in the preparation of novel epoxy
networks with improved re-resistance and tough-
ness. The temperature for a 5 % weight loss of the
modied epoxy networks occurred at around 400

C.
All phosphine-oxide-based epoxy thermosets yielded
16.717.7 % char at 750

C under air ow, whereas


the control sample, DGEBA/DDS, had no char at that
temperature.
Ito and Miyake
143
suggested the use of a commer-
cial monofunctional amide (Scheme 124), made by
reacting diethylphosphite with acrylamide. This was
efcient in epoxy resins heavily loaded with silica,
which have been used for the packaging of electronic
devices.
Phosphorus-containing guanamines (Schemes 125
and 126)
165
and a spirocyclic bisphosphonate (Scheme
1922 Polym Int 53:19011929 (2004)
Decomposition, combustion and re retardancy of epoxy resins
Scheme 125.
Scheme 126.
Scheme 127.
Scheme 128.
Scheme 129.
Scheme 130.
127)
129,166
were used to cure DGEBA resin. Printed
circuit board laminates containing 24wt% of the gua-
namines showed a V-1 rating, whereas those contain-
ing 33wt%of the spirocyclic bisphosphonate displayed
a V-0 rating.
The phosphonic acid made by reacting methylphos-
phorus dichloride with acrylic acid (Scheme 128),
167
or the related anhydride (Scheme 129)
168
were used
for the chain extension of DGEBA epoxy resins. Phos-
phonic diacids were also obtained by reacting the
anhydride with aliphatic diols (Scheme 130).
168
Epoxy
resins containing the anhydride or diacids cured with
DICY showed a V-0 rating.
Propylphosphonic anhydride (Scheme 131) was
used to cure epoxy resins composed of bisphenol
F and bisphenol A diglycidyl ethers and epoxy
novolac.
169
Hardened plates of 12 layers of glass
boards containing 3.5 % P showed good ame-
retardant performance in the DIN 5510-2 test.
Phosphonic diacids (Schemes 132 and 133) were used
to cure bisphenol F and novolac epoxy resins, which
provided V-0 ratings, in combination with ATH.
170
Scheme 131.
Scheme 132.
Scheme 133.
Scheme 134.
Scheme 135.
Scheme 136.
Scheme 137.
O-methyl P-ethyl-P-methylphosphinate (Scheme
134), dimethyl methylphosphonate (Scheme 135),
dimethyl ethylphosphonate (Scheme 136) or dimethyl
propylphosphonate (Scheme 137) were reacted with
P
2
O
5
to make pyrophosphinates or pyrophosphonates,
respectively, which were used to cure epoxies.
171
The bicyclic phosphate 1-oxo-4-hydroxymethyl-
2,6,7-trioxa-1-phosphabicyclo[2.2.2]octane (PEPA)
(Scheme 138) was combined with DGEBA epoxy
resin to obtain halogen-free ame-retardant poly-
mers.
172,173
These were compared to the resins cured
by ethylene diamine or phthalic anhydride. As mea-
sured by cone calorimetry, the ignition time increased
and the heat release rate, CO and smoke evolved,
decreased with PEPA content. During combustion,
the presence of intumescent char was observed, and
the char yield increased with increase of PEPA con-
tent. TGAand FTIRspectroscopy results demonstrate
an increase of crosslinking in the process of ther-
mal decomposition, which is partially responsible for
Polym Int 53:19011929 (2004) 1923
SV Levchik, ED Weil
Scheme 138.
Scheme 139.
Scheme 140.
Scheme 141.
Scheme 142.
charring. A monofunctional aliphatic phosphine oxide
(Scheme 139) was used in combination with ATH to
provide a V-0 rating to cast epoxies.
131
La Rosa et al
87
studied the ame-retardancy
of a dual-cure thermosetting system, consisting
of an epoxy resin blended with an unsaturated
polyester. Halogen-free ame-retarded formulations
were prepared by adding different amounts (5, 10 and
15 %) of three commercial phosphorus-containing
diols (Schemes 140142, respectively). According to
the OI numbers, no signicant variation between the
different ame-retardants was found.
Silicon-containing epoxy resins
Silica is widely used in molded epoxy formulations
for encapsulating electronic devices. Because of high
loading, it provides a ame-retardant effect mostly
due to the heat-sink effect. In contrast, silicon
incorporated in the epoxy resin network ensures a
chemical mode of the re retardant action, which is
observed at relatively low silicon contents.
Triglycidylphenyl silane (Scheme 143) was synthe-
sized and the corresponding silicon-containing epoxy
Scheme 143.
Scheme 144.
resins were prepared by curing with DDS, DICY
or DDM.
174
Fire-retardant and thermal decompo-
sition performances of the silicon-containing epoxy
were compared to those of commercial DGEBA epoxy
resins.
175
The introduction of silicon resulted in higher
curing reactivity. Char yields and OI measurements
demonstrated improved ame-retardancy, because the
OI increased from 24 for a standard commercial resin,
to 35 for silicon-containing resins.
In another study,
155
the same tri-functional
resin was combined with difunctional bis-(3-
glycidyloxy)phenylphosphosphine oxide (Scheme 12)
and cured either with a siloxane diamine (Scheme 144)
or bis(4-aminophenyl) phenylphosphonate (Scheme
100) curing agents. An OI enhancement from 26 to
36 was observed for epoxy resins containing both phos-
phorus and silicon. This synergistic effect was further
boosted by using siloxane. Epoxy resins consisting of
the phosphorus epoxide and siloxane diamine exhib-
ited a high OI value of 41. While under ame exposure,
phosphorus provided a tendency for char forma-
tion, silicon provided an enhancement of the thermal
stability of the char. Introducing both phosphorus
and silicon together in the epoxy resin composition
successfully combines these two factors in a strong
ame-retardation mechanism.
New silicon-containing epoxy monomers were
obtained by reacting DGEBA with diphenyl-
silanediol (Scheme 145) or o-cresol-formaldehyde
epoxy novolac with triphenylsilanol (Scheme 146).
176
These resins were cured with DOPO-containing
diamine, triphenylphosphine oxide diamine or DOPO-
phosphorylated phenol formaldehyde resin. With sil-
icon incorporation, the OI values of the epoxy resins
were increased from 21 to 22.5 and from 21 to 27 for
the silanediol (Scheme 145) and silanol (Scheme 146)
systems, respectively. When cured with phosphorus-
containing curing agents, the resulting epoxy resins
Scheme 145.
1924 Polym Int 53:19011929 (2004)
Decomposition, combustion and re retardancy of epoxy resins
Scheme 146.
Scheme 147.
Scheme 148.
Scheme 149.
Scheme 150.
showed extremely high OI values of 35 and 49 for
these two epoxy monomers (Schemes 145 and 146),
respectively.
A series of siloxane-containing diamines (Schemes
147149) were used to cure a dicyclopentadiene-
containing epoxy (Scheme 150).
177
Silicon-based
resins exhibited OI values of around 3134 while the
silicon-free epoxy resin cured with DDM exhibited an
OI value below 19.
A silicon-novolac hybrid resin was prepared by
reacting polydimethylsilane and phenol-formaldehyde
novolac;
178
25wt% of this silicon-containing resin
provided a V-0 rating in bisphenol A epoxy resin.
Various methoxy-phenyl or ethoxy-phenyl polysilox-
anes at 10wt% loadings showed V-0 ratings in
novolac epoxy resin cured with phenol-formaldehyde
novolac.
179
Methyl-phenyl silicone copolymerized
with some siloxane units showed V-0 ratings in a
similar resin at 9wt%.
180
The diglycidyl phenylphosphonate (Scheme 12)
was cured with DDM in the presence of tetraethoxysi-
lane (Scheme 151).
181
In the presence of the catalyst,
ie p-toluenesulfonic acid, tetraethoxysilane decom-
posed, which led to epoxy-silica hybrid materials with
a nanostructure via an in situ sol gel process. The
glass transition temperature of the hybrid epoxy resins
Scheme 151.
Scheme 152.
increased with silica content. The OI values of the
DGEBA-based epoxy resins increased from 26 to 31,
so demonstrating that introduction of silica into the
epoxy improved the ame-retardancy. The enhance-
ment of the ame retardancy with the increase of
silicon was not observed in the mechanically blended
silicapolymer systems; however, it was shown that
nanometer-scale silica could enhance polymer ame
retardancy. It was speculated that while being heated,
the silicon (in the form of silica) migrated to the sur-
face of the material because of the relatively lowsurface
potential energy of silicon. The silica on the surface
of the materials served as a heat barrier to protect the
inner layer of the polymers. This protecting effect was
observed in the TGA thermograms under air, where
silica greatly inhibited the oxidation weight loss of the
polymeric materials, and resulted in high char yields
for the polymers at temperatures higher than 700

C.
Similar organicinorganic hybrids were prepared
by Chiang and Ma
182
using tetraethoxysilane and
diethylphosphatopropyltriethoxysilane (Scheme 152).
DGEBA was modied by a coupling agent,
3-isocyanatopropyltriethoxysilane (Scheme 153), to
improve the compatibility of the organic and inor-
ganic phases. The char yield of pure epoxy resin was
14.8wt% while that of the modied epoxy nanocom-
posite was 31wt% at 800

C. Values for the oxygen


index of pure epoxy and the modied epoxy nanocom-
posites were 24 and 32, respectively.
Miscellaneous additives
Because of some difculties (or at least regulatory
impediments) with disposal, in particular the inciner-
ation of halogen-containing electronic devices, there
is a trend to search for suitable alternatives. Together
with phosphorus-containing products, aluminum tri-
hydrate (ATH) is nding applications in the ame-
retardancy of epoxy resins for printed wiring boards.
Various halogen-free formulations are often based
on high-performance epoxy with added ATH. These
early halogen-free systems found their niche in the
market placebut not only because of their environ-
mentally friendly characteristics. Some halogen-free
systems show a low coefcient of thermal expansion
and as a result, are more stable to delamination when
compared to TBBA-containing systems.
66
Polym Int 53:19011929 (2004) 1925
SV Levchik, ED Weil
Scheme 153.
Scheme 154.
A special grade of high thermally stable ATH
was made by heating freshly prepared highly dis-
persed ATH at 220

C for 2 h and then treat-


ing it with aminoalkylsilane.
183
A 70wt% load-
ing of this ATH provided a UL 94V-0 rating
in the printed circuit board laminates made with
DGEBA/DICY.
The effect of ATH and triphenyl phosphate (TPP)
on the ame-retardancy and thermal stability of
anhydride-cured epoxy resin was studied by Xiao
et al.
86
Surprisingly, it was found that reactive alumina
(a Lewis acid) was formed as the ATH lost water,
which could catalyze the degradation of the cured
epoxy resin. However, the co-operative use of ATH
and TPP could lead to higher char yield in spite of
the catalysis of degradation of the cured epoxy resin
and could display condensed-phase ame-retardation
synergism.
It has been illustrated that zinc borate alone can
outperform magnesium hydroxide and zinc carbonate
in halogen-free epoxy/novolac systems for electrical
applications.
184
Zinc borate could not only provide
a V-0 performance at 0.8 mm but also improved the
thermal stability and copper/epoxy adhesion.
Liu et al
147
studied the re-retardant perfor-
mance of cresol-novolac epoxy resins cured with
phenol-formaldehyde novolacs and novolac/DOPO-
containing copolymers. The epoxy resin exhibited high
glass transition temperatures (159177

C), high char


yields and high oxygen index values (2632.5), hence
indicating that the resins had good ame-retardance.
Using a melamine-modied phenol formaldehyde
novolac (Scheme 154) to replace the regular phe-
nol formaldehyde novolac in the curing composi-
tion further enhanced the cured epoxy resins glass
transition temperatures (160186

C) and oxygen
index values (2833.5). Preparation of the product
of condensation of melamine, phenol and resorci-
nol bis(diphenyl phosphate) with formaldehyde was
claimed by Honda et al.
185,186
A 28wt% of the con-
densate, used as a curing agent in combination with
23wt% of ATH, provided a V-0 rating for laminates
prepared with a blend (4:1) of DGEBA/cresol novolac
epoxy resin.
A similar effect of increasing OI was observed when
melamine-modied novolac was used to cure silicon-
containing epoxies.
176
However, char enhancement
was not observed for the melamine-containing resins.
It was speculated that under heat, the melamine
group transformed to melam and melem via a
deammoniation reaction. The ammonia gas and the
heat-resistant melam and melem could enhance the
OI through both gas-phase and condensed-phase
mechanisms.
CONCLUSIONS
The thermal stability of cured epoxy resins is
strongly affected by the curing agent used. Thermal
decomposition of any epoxy resin starts from the
dehydration of the secondary alcohol in the aliphatic
part of the resin, which leads to the formation
of vinylene ethers. The next step of thermal
decomposition is splitting of either the ether or amine
bonds, weakened by the allylic groups. Evolution of
highly combustible small aliphatic fragments follows.
Some of the aromatic structures left can undergo
charring. In the presence of ame-retardant additives,
the thermal decomposition accelerates, although the
amount of solid residue increases.
The combustion performance of epoxy resins is
strongly affected by the crosslinking density. In
general, epoxy resins cured with anhydrides are more
ammable than epoxy resins cured with amines. The
oxygen index tends to be positively correlated to the
char yield.
The ame-retardancy of epoxy resins can be
improved by using special epoxy monomers con-
taining highly aromatic bisphenols, eg phenolph-
thalein, bisphenol-uorenone, bisphenol-anthrone,
tetraphenol-anthracene or bisphenols containing dou-
ble bonds able to undergo the DielsAlder reaction
(eg mono-,di- or trihydroxystyrylpyridines). These
monomers, cured either alone or in combination with
regular-grade epoxy monomers, show a signicant
increase of char yield and oxygen index. Highly aro-
matic diamines used as curing agents also help to
improve the ame-retardancy. Some epoxy resins can
be copolymerized with other highly charrable resins,
which results in total improvement of the ame-
retardancy and physical properties.
For over 30years, tetrabromobisphenol A (TBBA)
has been used in epoxy resins as a co-reactant,
especially in electronic grade materials. Industry has
been largely satised with tetrabromobisphenol A and
therefore very limited activity has been observed in
the development of new halogen-containing ame
retardants for electronic applications. On the other
hand, some activity is observed in the composite
materials where high charring tendency and lowsmoke
formation are in high demand.
Phosphorus-containing re-retardants are partic-
ularly effective in epoxy resins because the phos-
phate esters or phosphoric acids formed in the ther-
mal decomposition of phosphorus-containing prod-
ucts tend to react with OH groups and involve
the epoxy resin structure in the charring. Various
1926 Polym Int 53:19011929 (2004)
Decomposition, combustion and re retardancy of epoxy resins
phosphorus-containing additives, eg red phosphorus,
ammonium polyphosphate, salts of alkylphosphinic
or alkylphosphonic acids, phosphonate esters and
phosphate esters, have been shown to be effective
ame-retardants in epoxy resins. Depending on the
structure of the epoxy resin, curing agent and pres-
ence of ller, from 2 to 5 wt% phosphorus is required
in order to achieve a V-0 rating in the UL-94 test.
Phosphorus-containing additives often decrease the
physical properties of the epoxy resin; in particular,
phosphate esters tend to decrease the glass transition
temperature.
There is a relatively large number of publica-
tions regarding reactive phosphorus-containing epoxy
monomers. Special attention has been paid to the use
of 9,10-dihydro-9-oxa-10-phosphaphenanthrene 10-
oxide (DOPO). The latter can be either pre-reacted
with the epoxy resin or used as a reactive additive
during curing of the epoxy resin. Hydroquinone-
or naphthalene-type diphenols with a DOPO sub-
stituent were used in the chain extension of various
epoxy resins. Because of the highly aromatic struc-
ture of these epoxies, a relatively low content of
phosphorus provides a V-0 rating and high OI val-
ues. Bis-glycidyloxy aryl- or alkylphosphonates were
extensively studied as comonomers with regular epoxy
resins. The phenylphosphonate structure was used as
a building block in various epoxy resins because of
the high efciency and high thermal stability of this
structural unit. Aromatic phosphate groups were also
incorporated into epoxy monomers; however, hin-
dered structures were required in order to provide
sufcient hydrolytic stability.
In a similar way to epoxy monomers, DOPO was
extensively studied as a phosphorus-containing group
in various curing agents. Unmodied DOPO was
recommended as a part of the curing system for
halogen-free ame-retardant epoxies and provided
a UL 94V-0 rating with phosphorus contents
below 3 %. Similar to the phosphorus-containing
monomers, phenylphosphonate or aromatic hindered
phosphate groups were found suitable in curing
agents; however, special attention in the literature
has been paid to triphenylphosphine oxide or
diphenylalkylphosphine oxide structures. Phosphine
oxides, in particular, are extremely hydrolytically and
thermally stable. Phosphoramides were shown to be
reactive with epoxy resins, and therefore have been
considered as potential phosphorus-containing curing
agents. A combination of phosphorus-containing
epoxy monomers with phosphorus-containing curing
agents gave a particularly good ame-retardant effect
with an OI as high as 50.
Silicone has been incorporated in both the epoxy
monomers and in the curing agents. This was found
to be particularly effective when in combination
with phosphorus. A melamine-phenol-formaldehyde
condensate provided a curing agent rich in nitrogen
which was shown to be synergistic in phosphorus-
containing epoxies.
REFERENCES
1 McAdams LV and Gannon JA, Encyclopedia of Polymer Science
and Engineering, Vol 6, Wiley, New York, pp 322382
(1988).
2 Epoxy Resins, Chemistry and Technology, ed by May CA, Marcel
Dekker, New York (1988).
3 Chemistry and Technology of Epoxy Resins, ed by Ellis B, Blackie
Academic and Professional, London (1993).
4 Dyakonov T, Mann PJ, Chen Y and Stevenson WTK, Polym
Degrad Stabil 54:67 (1996).
5 Iji M and Kiuchi Y, Polym Adv Technol 12:393 (2001).
6 Puglia D, Manfredi LB, Vazquez Aand Kenny J, PolymDegrad
Stabil 73:521 (2001).
7 Christiansen W, Shirrell D, Aguirre B and Wilkins J, in
Proceedings of IPC Printed Circuits EXPO 2001, Anaheim,
CA, USA, pp S03-1-1S03-1-7 (2001).
8 Lee L-H, J Appl Polym Sci 9:1981 (1965).
9 Lee L-H, J Polym Sci Part A 3:859 (1965).
10 Keenan MA and Smith DA, J Appl Polym Sci 11:1009 (1967).
11 Bishop DP and Smith DA, Ind Eng Chem (August):32 (1967).
12 Paterson-Jones JC and Smith DA, J Appl Polym Sci 12:1601
(1968).
13 Leisegang EC, Stephen AM and Paterson-Jones JC, J Appl
Polym Sci 14:1961 (1970).
14 Paterson-Jones JC, Percy VA, Giles RGF and Stephen AM, J
Appl Polym Sci 17:1867 (1973).
15 Paterson-Jones JC, Percy VA, Giles RGF and Stephen AM, J
Appl Polym Sci 17:1877 (1973).
16 Paterson-Jones JC, J Appl Polym Sci 19:1539 (1975).
17 Chen CS, Bulkin BJ and Pearce EM, J Appl Polym Sci 28:1077
(1983).
18 Maxwell ID and Pethrick RA, Polym Degard Stabil 5:275
(1983).
19 Grassie N, Guy MI and Tennent NH, Polym Degrad Stabil
14:125 (1986).
20 Levchik SV, Camino G, Luda MP, Costa L, Costes B,
Henry Y, Morel E and Muller G, Polym Adv Technol 6:53
(1995).
21 Levchik SV, Camino G, Luda MP, Costa L, Costes B,
Henry Y, Muller Gand Morel E, PolymDegrad Stabil 48:359
(1995).
22 Bishop DP and Smith DA, J Appl Polym Sci 14:205 (1970).
23 Neiman MB, Kovarskaya BM, Golubenkova LI, Strizhkova
AS, Levantovskaya II and Akutin MS, J Polym Sci 56:383
(1962).
24 Stuart JM and Smith DA, J Appl Polym Sci 9:3195 (1965).
25 Becker L, Lenoir D, Matuschek G and Kettrup A, J Anal Appl
Pyrol 60:55 (2001).
26 Lin SC, Bulkin BJ and Pearce EM, J Polym Sci Polym Chem Ed
17:3121 (1979).
27 Dante MF and Conley RT, Am Chem Soc Div Org Coatings
Plast Chem Prepr 24:135 (1964).
28 Cerceo E, Ind Eng Chem Prod Res Devel 9:96 (1970).
29 Le Huy HM, Bellenger V, Verdu J and Paris M, Polym Degrad
Stabil 35:77 (1992).
30 Le Huy HM, Bellenger V, Paris M and Verdu J, Polym Degrad
Stabil 35:171 (1992).
31 Burton BL, J Appl Polym Sci 47:1821 (1993).
32 Rose N, Le Bras M, Delobel R, Costes B and Henry Y, Polym
Degrad Stabil 42:307 (1993).
33 Rose N, Le Bras M, Bourbigot S and Delobel R, Polym Degrad
Stabil 45:387 (1994).
34 Rose N, Le Bras M, Bourbigot S, Delobel R and Costes B,
Polym Degrad Stabil 54:355 (1996).
35 Le Bras M, Rose N and Bourbigot S, J Fire Sci 14:199 (1996).
36 Lesnikovich AI and Levchik SV, J Therm Anal 27:89 (1983).
37 Sanchez G, Brito Z, Mujica V and Perdomo G, Polym Degrad
Stabil 40:109 (1993).
38 Luda MP, Balabanovich AI and Camino G, J Anal Appl Pyrol
65:25 (2002).
39 Liu Y-L, Hsiue G-H, Lan C-W and Chiule Y-S, Polym Degrad
Stabil 56:291 (1997).
Polym Int 53:19011929 (2004) 1927
SV Levchik, ED Weil
40 Wu CS, Liu YL, Chiu YC and Chiu YS, Polym Degrad Stabil
78:41 (2002).
41 Chen CS, Bulkin BJ and Pearce EM, J Appl Polym Sci 27:1177
(1982).
42 Yan HJ and Pearce EM, J Appl Polym Sci 22:3319 (1984).
43 Yang C-P and Lee T-M, J Polym Sci Polym Chem Ed 28:887
(1990).
44 Macaione DP, Dowling RP, II and Bergquist PP, AMMRCTR
8353, Report of Army Materials and Mechanics Research
Center, Watertown, MA, USA (1983).
45 Iji M and Kiuchi Y, J Mater Sci Mater Electron 12:715 (2001).
46 Martin FJ and Price KR, J Appl Polym Sci 12:143 (1968).
47 Lin SC and Pearce EM, J Polym Sci Polym Chem Ed 17:3095
(1979).
48 Chen CS, Bulkin BJ and Pearce EM, J Appl Polym Sci 27:3289
(1982).
49 Hshieh F-Y and Benson HD, Fire Mater 21:41 (1997).
50 Arada B, Lin SC and Pearce EM, Int J Polym Mater 7:167
(1979).
51 Kiuchi Y and Iji M, in Proceedings of Conference on Recent
Advances in Flame Retardancy of Polymeric Materials,
Stamford, CT USA, pp 280287 (2001).
52 Shobara T, Okuse S, Aoki T and Kato H (to Shin-Etsu), US
Patent 6 143 423 (2000).
53 Maeda M and Iwasaki S (to Sumitomo Bakelite), US Patent 6
190 787 (2001).
54 Iwasaki S, Iji M and Kiuchi Y (to Sumitomo Bakelite and
NEC) US Patent 6 242 110 (2001).
55 Wang T-S, Yeh J-F and Shau M-D, J Appl Polym Sci 59:215
(1996).
56 Chin W-K, Shau M-D and Tsai W-C, J Polym Sci Polym Chem
Ed 33:373 (1995).
57 Hwang S-H and Jung J-C, J Appl Polym Sci 81:279 (2001).
58 Manfredi LB, Claro JA, Kenny JM, Mondragon Egana I and
Vazquez A, Polym Compos 20:675 (1999).
59 Tyberg CS, Lin SL, Bergeron K, Sankarapandian M, Shih P,
Loos AC, McGrath JE and Rife JS, in Proceedings of the 44th
International SAMPE Symposium, Long Beach, CA, USA, pp
12061216 (1999).
60 Tyberg CS, Shih P, Verghese KNE, Loos AC, Lesko JJ and
Rife JS, Polymer 41:9033 (2000).
61 Rimdusit S and Ishida H, in Proceedings of ANTEC99
Conference, New York, pp 935939 (1999).
62 Additive Enhances Thermoset Resins, Reinforced Plast (May):
20 (2000).
63 Murai H, Takeda Y, Takano N and Ikeda K, Circuit World,
27:7 (2001).
64 IPC White Paper on Halogen-Free Materials Used for Printed
Circuit Boards and Assemblies, Draft Three, IPC, Northbrook,
IL, USA (2001).
65 Brown N and Aggleton M, Reinforced Plast (October): 44
(1999).
66 Demaree R, in Proceedings of IPC Printed Circuits EXPO 2001,
Anaheim, CA, USA, pp S11-3-1S11-3-4 (2001).
67 Arata M, Sase S, Takano N and Takeda Y (to Hitachi), US
Patent 6 180 250 (2001).
68 Wang C-S, Berman JR, Walker LL and Mendoza A, J Appl
Polym Sci 43:1315 (1991).
69 Markezich RL, in Proceedings of FRCAConference, Philadelphia,
PA, USA, pp 113126 (2001).
70 Lo J and Pearce EM, J Polym Sci Polym Chem Ed 22:1707
(1984).
71 Lyon RE, Castelli LM and Walters RN, in Proceedings of
Conference on Recent Advances in Flame Retardancy of
Polymeric Materials, Stamford, CT, USA, pp 267279
(2001).
72 Morgan RJ, in Advances in Polymer Science, Vol 72, ed
by Dusek K, Springer-Verlag, Berlin, Germany, pp 343
(1985).
73 Spennger S and Utz R, in Proceedings of Conference Addcon
Word98, London, paper no.18 (1998).
74 Spennger S and Utz R, J Adv Mater 33:24 (2001).
75 Jain P, Choudhary V and Varma IK, J Macromol Sci Polym Rev
C42:139 (2002).
76 Honda N and Sugiama T (to Toshiba) US Patent 5 994 429
(1999).
77 Osada S, Asano E, Ino S, Aoki T, Tomiyoshi K and Shio-
bara T (to Shin-Etsu), US Patent 6291 556 (2001).
78 Hoerold S, Walz R and Zopes H-P, Reinforced Plast (January):
40 (2000).
79 Levchik SV, Levchik GF, Antonov AV, Yablokova MYu,
Tuzhikov OI and Tuzhikov OO, in Proceedings of Confer-
ence on Recent Advances in Flame Retardancy of Polymeric
Materials, Stamford, CT, USA, pp 161176 (1999).
80 Choate M, in Proceedings of IPC Printed Circuits EXPO 2001,
Anaheim, CA, USA, pp S11-2-1S11-2-5 (2001).
81 Hoerold S (to Clariant), US Patent 6 420 459 (2002).
82 Iwata M, Seki N, Inoue K, Takahashi R, Fukumura T and
Tanaka M (to Chisso), US Patent 5 945 467 (1999).
83 Kodolov VI, Shuklin SG, Kuznetsov AP, Makarova IC, Bys-
trov SC, Demicheva OV and Rudakova TA, J Appl Polym
Sci 85:1477 (2002).
84 Zubkova NI, Butylkina NG, Khalturinski NA and Berlin AA
(to Isle Firestop), PCT Patent Application WO 00/14094
(2000).
85 Zubkova NI, Butylkina NG, Khalturinski NA, Berlin AA,
Vilesova MS, Bosenko MS and Voronkova LI, (to Isle
Firestop), PCT Patent Application WO 00/14094 (2000).
86 Xiao W, Xe P, Hu G and He B, J Fire Sci 19:369 (2001).
87 La Rosa AD, Recca A, Carter JT and McGrail PT, Polymer
40:4093 (1999).
88 Ito M, Miyake S, Shibata K and Tobisawa A (to Sumitomo
Bakelite), US Patent 5 932 637 (1999).
89 Derouet D, Morgan F and Brosse C, J Appl Polym Sci 62:1855
(1996).
90 Tobisawa A (to Sumitomo Bakelite), European Patent Applica-
tion 1 116 774 (2001).
91 Liu Y-I, Pearce EM and Weil ED, J Fire Sci 17:240 (1999).
92 Shimizu K, Tokunaga A and Tanaka M (to Toray), US Patent
5 919 844 (1999).
93 Tachikawa H, Takachika K, Saito S and Nagayama N (to
Asahi), US Patent 6 348 523 (2002).
94 Zerda AS and Lesser AJ, J Appl Polym Sci 84:302 (2002).
95 Dellar RJ, Richardson N and Clubley BG (to Ciba-Geigy), US
Patent 4 918 122 (1990).
96 Richardson J and Dellar RJ (to Ciba-Geigy), US Patent 4 972
011 (1990).
97 Richardson J and Dellar RJ (to Ciba-Geigy), European Patent
0 245 207 (1992).
98 Gan J, Goodson A, Koenig RA and Everett JP (to Dow
Chemicals), PCT Patent Application WO 99/00451 (1999).
99 Archer AC, Harris CJ, Woodward G and Zakikhani M (to
Rhodia), European Patent 0 758 654 (2001).
100 Denq B-L, Hu Y-S, Chen L-W, Chiu W-Y and Wu T-R, J
Appl Polym Sci 74:229 (1999).
101 Mirkamilov TM and Mukhamedgaliev BA, Int Polym Sci
Technol 28:T/40 (2001).
102 Li J, Chen S and Xu X, J Appl Polym Sci 40:417 (1990).
103 Maltseva EV, Rudakova TA, Demicheva OV, Khalturin-
ski NA, Panova LG and Artemenko SE, Int J Polym Mater
47:95 (2000).
104 Von Gentzkow W, in Proceedings of Conference on Recent
Advances in Flame Retardancy of Polymeric Materials,
Stanford, CT, USA, pp 215223 (1996).
105 Utz Rand Sprenger S (to Schill and Seilacher), European Patent
Application 0 806 429 (1997).
106 Gan J and Goodson A (to Dow Chemicals), PCT Patent
Application WO 01/42359 (2001).
107 Wang C-S and Lin CH, J Polym Sci Polym Chem Ed 37:3903
(1999).
108 Wang C-S and Lin CH (to National Science Council of
Taiwan), US Patent 6 291 627 (2001).
109 Lin CH and C-S Wang, Polymer 42:1869 (2001).
110 Lin CH and Wang CS, Am Chem Soc Div Polym Chem Polym
Prepr 43:908 (2002).
1928 Polym Int 53:19011929 (2004)
Decomposition, combustion and re retardancy of epoxy resins
111 Wang C-S and Shieh J-Y, J Appl Polym Sci 73:353 (1999).
112 Wang C-S and Lin CH, J Appl Polym Sci 75:429 (2000).
113 Wang C-S and Shieh J-Y, Polymer, 39:5819 (1998).
114 Shieh J-Y and Wang C-S, J Polym Sci Polym Chem Ed 40:369
(2002).
115 Asana T, Ogasawa K and Ito N (to Matsushita), European
Patent Application 1 103 575 (2001).
116 Wang C-S and Lin CH (to National Science Council of
Taiwan), US Patent 6 291 626 (2001).
117 Wang C-S and Lee M-C, Polymer 41:3631 (2000).
118 Lin CH, Wu CY and Wang C-S, J Appl Polym Sci 78:228
(2000).
119 Hoerold S and Kleiner H-J (to Clariant), US Patent 5 959 043
(1999).
120 Alcon MJ, Espinosa MA, Galia M and Cadiz V, Macromol
Rapid Commun 22:1265 (2001).
121 Liu Y-L, J Appl Polym Sci 83:1697 (2002).
122 Liu Y-L, J Polym Sci Polym Chem 40:359 (2002).
123 Chiu Y-S, Jiang M-D and Liu Y-L (to Chung-Chan Institute
of Science and Technology, Taiwan), US Patent 6 441 067
(2002).
124 Cheng K-C, Yu S-Y and Chiu W-Y, J Appl Polym Sci 83:2741
(2002).
125 Liu Y-L, Hsiue G-H, Chiu Y-S and Jeng R-J, J Appl Polym Sci
61:1789 (1996).
126 Cheng K-C, Yu S-Y and Chiu W-Y, J Appl Polym Sci 83:2733
(2002).
127 Liu Y-L, Hsiue G-H and Chiu Y-S, J Polym Sci Polym Chem
Ed 35:565 (1997).
128 Schutyser JAJ, Buser AJWand Steenbergen A(to Akzo Nobel),
PCT Patent Application 93/11176 (1993).
129 Von Gentzkow W, Schmidt E, Slavk M and Jurgen H-J (to
Siemens), German Patent Application DE 198 56 397 (2000).
130 Von Gentzkow W, Heinl D, Kapitza H and Schreyer M (to
Siemens), US Patent 6 201 074 (2001).
131 Wipfelder E and Plundrich W (to Siemens), European Patent 0
412 425 (1996).
132 Bhuniya SP and Maiti S, J Ind Chem Soc 77:482 (2000).
133 Shau M-D and T-S Wang, J Polym Sci Polym Chem Ed 34:387
(1996).
134 Wang T-S, Parng J-K and Shau M-D, J Appl Polym Sci 74:413
(1999).
135 Zigon M and Malavasic T, Kovine zlitine tehnologijie 31:227
(1997).
136 Von Gentzkow W, Huber J, Kapitza H, Rogler W, Kleiner H-J
and U Schonamsgruber (to Hoechst), US Patent 5 756 638
(1998).
137 Von Gentzkow W, Huber J, Kapitza H and Rogler W (to
Siemens), US Patent 5 760 146 (1998).
138 Scholz G, Hoerold S and Ping W-D (to Clariant), European
Patent 0 794 206 (2002).
139 Jeng R-J, Shau S-M, Lin J-J, Su W-Cand Chiu Y-S, Eur Polym
J 38:683 (2002).
140 Chen-Yang YW, Lee HF and Yuan CY, J Polym Sci Polym
Chem Ed 38:972 (2000).
141 Lengsfeld H, Altstadt V, Sprenger S and Utz R, Kunststoffe
91:37 (2001).
142 Cho C-S, Fu S-C, Chen L-W and Wu T-R, Polym Int 47:203
(1998).
143 Ito M and Miyake S (to Sumitomo Bakelite), US Patent 6 180
695 (2001).
144 Ito M and Miyake S (to Sumitomo Bakelite), US Patent 6 296
940 (2001).
145 Liu YL, Polymer 42:3445 (2001).
146 Shieh J-Y and Wang C-S, J Appl Polym Sci 78:1636 (2000).
147 Liu Y-L, Wu CS, Hsu KY and Chang TC, J Polym Sci Polym
Chem Ed 40:2329 (2002).
148 Wang C-S and Lin C-H, J Appl Polym Sci 74:1635 (1998).
149 Varma ID and Gupta U, J Macromol Sci Chem A23:19 (1986).
150 Levchik SV, Camino G, Luda MP, Costa L, Muller G,
Costes B and Henry Y, Polym Adv Technol 7:823 (1996).
151 Levchik SV, Camino G, Costa L and Luda MP, Polym Degrad
Stabil 54:317 (1996).
152 Levchik SV, Camino G, Luda MP, Costa L, Muller G and
Costes B, Polym Degrad Stabil 60:169 (1998).
153 Tchatchoua C, Ji Q, Srinivasan SA, Ghassemi H, Yoon TH,
Martinez-Nunez M, Kashtwagi T and McGrath JE, Am
Chem Soc Div Polym Chem Polym Prepr 38:113 (1997).
154 Liu Y-L, Hsiue G-H, Lee R-H and Chiu Y-S, J Appl Polym Sci
63:895 (1997).
155 Hsiue G-H, Liu Y-L and Tsiao J, J Appl PolymSci 78:1 (2000).
156 Wang C-S and Shieh J-Y, Eur Polym J 36:443 (2000).
157 Bright DA and Telschow JE (to Akzo Nobel), US Patent 6 403
819 (2002).
158 Brennan DJ, Everett JP and Nader B (to Dow Chemicals), US
Patent 6 403 220 (2002).
159 Jeong KU, Park IY, Kim IC and Yoon TH, J Appl Polym Sci
30:1198 (2001).
160 Asano E, Aoki T, Shiobara T, Flury P, Scharf W and Okada T
(to Shin-Etsu Chemicals and Ciba Specialty), European
Patent 0 742 261 (2000).
161 Kuo P-L, Wang J-S, Chen P-C and Chen L-W, Macromol
Chem Phys 202:2175 (2001).
162 Bertram JL and Davis W (to Dow Chemical), US Patent 4 481
347 (1984).
163 Jeng R-J, Wang J-R, Lin J-J, Liu Y-L, Chiu Y-S and Su W-C,
J Appl Polym Sci 32:3526 (2001).
164 Liu YN, Ji Q and McGrath JE, Am Chem Soc Div Polym Chem
Polym Prepr 38:223 (1997).
165 Buser AJW and Schutyser JAJ (to Akzo Nobel), US Patent 5
821 317 (1998).
166 Von Gentzkow W, Schmidt E, Slavk M and Jurgen H-J (to
Siemens), German Patent Application 19 856 396 (2000).
167 Hoerold S (to Clariant), German Patent Application 19 613 067
(1998).
168 Hoerold S and Schmitz H-P (to Clariant), European Patent 0
799 848 (2001).
169 Plundrich W and Wipperder E (to Siemens), PCT Patent
Application, WO 99/45061 (1999).
170 Wipfelder E and Plundrich W (to Siemens), US Patent 5 811
486 (1998).
171 Hoerold S and Scholz G (to Hoechst), European Patent
Application 0 794 205 (1997).
172 Ou Y, Li X and Peng Z, in Proceedings of FRCA Conference,
San Francisco, CA, USA, pp 179190 (2001).
173 Li X, Ou Y and Shi Y, Polym Degrad Stabil 77:383 (2002).
174 Hsiue G-H, Wang W-J and Chang F-C, J Appl Polym Sci
73:1231 (1999).
175 Wang WJ, Peng LH, Hsiue GH and Chang FC, Polymer
41:6113 (2000).
176 Wu CS, Ling YL and Chiu YS, Polymer 43:4277 (2002).
177 Hsiue G-H, Wei H-F, Shiao S-J, Kuo W-J and Sha Y-A, Polym
Degrad Stabil 73:309 (2001).
178 Lee M-S, Kang C-J and Peng K-L (to Industrial Technology
Research Institute), US Patent 6 337 363 (2002).
179 Yamamoto K, Yamaya M, Yamamoto A and Kobayashi Y (to
Shin Etsu), US Patent 6 184 312 (2001).
180 Itagaki A, Yamaya M and Kobayashi Y (to Shin-Etsu), US
Patent 6 326 425 (2001).
181 Hsiue G-H, Liu Y-L and Liao HH, J Polym Sci Polym Chem
Ed 39:986 (2001).
182 Chiang C-L and Ma C-CM, Eur Polym J 38:2219 (2002).
183 Brown N and Agglenton M (to Alusuisse Martinswerk), US
Patent 6 280 839 (2001).
184 Shen KK, in Proceeding of Conference on Recent Advances in
Flame Retardancy of Polymeric Materials, Stamford, CT, USA,
pp 7583 (2001).
185 N. Honda, T. Sugiama and T. Suzuki, (to Toshiba), US Patent
5 955 184 (1999).
186 N. Honda, T. Sugiama and T. Suzuki, (to Toshiba), US Patent
6 214 455 (2001).
Polym Int 53:19011929 (2004) 1929

Das könnte Ihnen auch gefallen