Vous êtes sur la page 1sur 25

Bismaleimides

Related terms:

Viscosity, Monomers, Oligomers, Phenolics, resin, Bismaleimide, Cyanate Ester,


Higher Temperature

View all Topics

Learn more about Bismaleimides

Step Polymerization
Bernard Sillion, in Comprehensive Polymer Science and Supplements, 1989

30.4.2.1(iv) Cross-linking of bismaleimides


When bismaleimides are heated to 210 °C, they polymerize by an addition mecha-
nism with the formation of a cross-linked network (76; equation 16).135

(16)

In the case of rigid bismaleimides with high melting points, the correlation between
structure and polymerization temperature shows that polymerization has to be
carried out above 230 °C.136 When R is a diphenyl sulfone or a diphenyl ether sulfone
oligomer (77), the more flexible linkages make it possible to obtain a B-stage by
heating to 180–200 °C. A postcure at 270 °C gives the final network by a radical
polymerization mechanism.137

77
When R is aliphatic, it was shown by IR spectroscopy that, up to a conversion of
20–30%, polymerization was pseudo first order. There is a relationship between
the length of the methylene sequence (6 < n < 12) and the activation energy,
which decreases with decreasing n.123 Bismaleimides can also be polymerized and
cross-linked with a basic catalyst such as 2-methylimidazole or diazabicyclooctane,
by an anionic mechanism.138

> Read full chapter

Thermoset Resins: The Glue That


Holds the Strings Together
F.C. Campbell, in Manufacturing Processes for Advanced Composites, 2004

3.4 Bismaleimide Resins


Bismaleimides were developed to bridge the temperature gap between epoxies and
polyimides with dry Tg's in the range of 430–600 °F and usage temperatures of
300–480 °F. They process similar to epoxies, by curing through addition reactions
at 350 °F. To obtain their high-temperature properties, they are given free-standing
post-cures at 450–475 °F to complete the polymerization reactions. BMI composites
can be processed by autoclave curing, filament winding and resin transfer molding.
The tack and drape of most BMIs are quite good due to the liquid component of the
reactants. Since BMIs process at the same temperature (i.e., 350 °F) and pressures
(e.g., 100 psig) as epoxies, conventional nylon bagging films, bleeder and breather
materials and other expendables can be used. In contrast, the higher-temperature
traditional polyimide materials usually require higher-temperature cures (600–700
°F) and usually higher pressures (e.g., 200 psig or higher) resulting in more expensive
and difficult tooling and bagging materials.

BMI chemistry is quite varied, with many potential paths to producing matrix mate-
rials. Both bismaleimides and polyimides contain the imide group shown in Fig. 16.
Bismaleimide monomers are synthesized by reactions with a primary diamine with
maleic anhydride (Fig. 17). The most prevalent BMI base monomer for matrices and
adhesives is 4,4 -bismaleimidedodiphenylmethane.
Fig. 16. Bismaleimide Chemical Structure

Fig. 17. Synthesis of BMI Resin

The wide range of possible chemical reactions involved in curing BMIs are shown in
Fig. 18. Homopolymerization (Fig. 18a) under heat is simple, but results in extremely
brittle structures. A more practical approach (Fig. 18b) is copolymerization with a
diamine, which produces a chain of more extended and flexible polymer. Copoly-
merization with an olefinic compound is another route to toughening, as shown for
divinylbenzene (Fig. 18c), bispropenylphenoxy (Fig. 18d) and diallylphenyl or dial-
lyphenoxy compounds (Fig. 18e). All three of these approaches have been successful-
ly used in commercial products such as V378A, Matrimid 5292 and Compimide 796.
BMIs have also been reacted with epoxies (Fig. 18f) in attempts to provide epoxy-like
processing with BMI high-temperature performance. The copolymerization with
0,0 -discyanobisphenol A (Fig. 18g) has been used to produce some resins known
as BT resins, where the B stands for BMI and the T stands for triazine, a type of
cyanate ester. The BCB (benzocyclobutene)-BMI copolymers (Fig. 18h) exhibit high
Tg's, good thermal-oxidative behavior, high toughness and processing ease, but the
BDB monomer is expensive and commercially unavailable. The ATT (addition type
thermoplastics), in which certain acetylene-terminated materials (Fig. 18j) undergo
copolymerization with BMIs, allows the use of addition curing low-molecular-weight
and low-viscosity starting materials. In addition, due to their linear backbone they
are inherently tough but also possess good thermal-oxidative stability as a result of
a fully aromatic structure.
Fig. 18. Chemistry of BMI Resins

Commercial BMIs are usually one of five forms:8 (1) BMIs or mixtures of different
BMIs, (2) blends of BMIs and BMI-diamines, (3) BMI and olefinic monomers and/or
oligomelic blends, (4) BMI and epoxy blends and (5) BMI and 0,0 -dicyanobisphenol
A mixtures.

Although early BMI materials were characterized as being hard to process (low tack
and short out-times) and possessing low toughness (brittle), BMI materials being
produced today have much better tack and long out-times. In addition, some BMIs
(e.g., Cytec 5250–4) are as tough as most toughened epoxies. Using liquid molding
processes such as RTM can also readily process bismaleimides.

One potential usage problem with bismaleimides and any polymer containing the
imide end-group is a phenomenon known as “imide corrosion.” This is a form of
hydrolysis that results in degradation of the polymer itself. It was originally observed
with carbon fiber/bismaleimide composites, galvanically coupled to aluminum in
a sump environment (i.e., a mixture of salt water and jet fuel). If the aluminum
corrodes, the composite (since it is electrically coupled to the aluminum through
the carbon fibers) becomes the cathode. Water reduction in the presence of oxygen
occurs at the cathode leading to the formation of hydroxyl ions that attack the
imide-carbonyl linkage of the BMI. The mechanism is shown schematically in Fig.
19. No corrosion occurs with non-conductive fibers such as glass or aramid, nor
does corrosion occur if the metal is galvanically similar to carbon, such as titanium or
stainless steel. Studies9 have shown that increases in temperature and bare exposed
carbon edges accelerate the deterioration. Electrically isolating the carbon from
the aluminum prevents the problem. This can be accomplished by curing a layer
of fiberglass on the composite faying surface and then sealing the edges with a
polysulfide sealant.
Fig. 19. Galvanic Corrosion Mechanism for Imide Linkage

> Read full chapter

Assessing the moisture resistance of ad-


hesives for marine environments
W. Broughton, in Adhesives in Marine Engineering, 2012

7.2.5 Polyimides and bismaleimides


Polyimides and bismaleimides are relatively expensive structural adhesives that
have excellent long-term strength retention at elevated temperatures, suitable for
prolonged use at temperatures between 200 °C and 350 °C. Poly imide and bis-
maleimide adhesives also have good chemical and moisture resistance. Polyimides
can perform up to 260 °C, exhibiting superior bond strength to epoxy-phenolics
at this temperature. The presence of moisture does have a detrimental effect at
elevated temperatures with marked bond strength reduction at temperatures above
260 °C. Bismaleimides are purported to have very high resistance to moisture even at
elevated temperatures (200 °C to 350 °C), although as with other adhesive systems,
environmental resistance decreases at elevated temperatures. Disadvantages for
both adhesive types include difficulty in handling, high curing temperatures, long
cure times and innate brittleness.

> Read full chapter

Adhesives and Sealants


Guy Rabilloud, in Handbook of Adhesives and Sealants, 2006
4.5.3.1 Maleimide-Terminated Adhesives
As previously stated, bismaleimides are crystalline compounds with melting points
in the range 100–210°C. A few examples are given in Table 4 which provides
the melting temperatures Tm, the maximum temperatures of the exothermic
peak Tmax and the polymerisation enthalpies ΔH. Diamines based on 4,4 -meth-
ylenebisbenzeneamine carrying methyl and ethyl substituents exhibit melting
points around 250°C, while the only bismaleimide melting at about 100°C is the
compound obtained from the reaction of maleic anhydride with 6-amino-1-(4-
phenyl)-1,3,3-trimethylindane.

Table 4. Melting temperature Tm, temperature at the maximum of the exothermic


peak Tmax and polymerisation enthalpy ΔH for some bismaleimides

Ar Tm(°C) Tmax(°C) ΔH ( J g− 1)

155–157 235 198


150–154 298 187
149–151 328 206
235 290 216
90–100 203 89

The chemistry of bismaleimides has been briefly discussed in Section 4.3.4.1 which


outlines that neat BMI resins form brittle networks. The values characterising the
shock resistance and crack propagation of a given material are the stress intensity
factor, K1C, and the fracture energy, G1C, which indicates the relaxation of the strain
energy. In highly cross-linked networks, such as those formed by tetrafunctional
epoxy resins, K1C is of the order of 0.5 MPa m0·5 and G1C values are around 60 J m− 2,
whereas the data obtained for a thermoplastic polyimide are respectively 2.8 MPa m-
0.5 and 3.5 kJ m− 2. The fracture toughness of cured BMIs is approximately 100 J m− 2,

increasing by a factor of two on chain extension between the two terminal double
bonds. For example, the fracture energy of polymer 40 (Fig. 17) is enhanced when
the molecular ratio of BMI 38 to 4,4 -methylenebisbenzeneamine 34 is increased.

Several methods have been investigated to improve the fracture toughness of BMI
resins, the most used technique being the incorporation of a second phase formed
by the dispersion of rubber particles into the cross-linked matrix. Carboxyl-termi-
nated butadiene-acrylonitrile elastomers (CTBN), which are very effective with epoxy
resins, are also frequently added to BMI adhesives. The general formula of CTBN
rubbers is:

where the number average molecular weight is in the range of 1300–3500 g mol− 1


with, in this later case, typical values of 5 and 10 for m and n, respectively. Starting
from a homogeneous mixture, three steps characterise the cure cycle: phase segre-
gation, gelation and vitrification. The final material is composed of rubber particles
with average diameters between 1 and 5 μm dispersed within the continuous BMI
matrix. The maximum fracture resistance is obtained for 15–20% CTBN, resulting
in fracture energy values multiplied by a factor of 20 to 30. It has been demonstrated
that mixtures of CTBN Hycar 1300-X8 and bismaleimide Compimide 353 could be
copolymerised at 170–200°C to provide two-phase materials containing dispersed
rubber microspheres [69]. Fig. 25 shows that the fracture energy increases with the
quantity of CTBN rubber, which can be as high as 50%. The flexural modulus at
high temperature (250°C), however, rapidly decreases to become close to zero at
50% CTBN concentration.

Figure 25. Changes of the fracture energy (a) and flexural modulus at 25°C (b) and
250°C (c) for CTBN-BMI adhesives as a function of the CTBN-rubber content [69].

Another method used to improve the fracture energy of BMI resins consists in
mixing the thermosetting material with linear thermoplastic polymers. This can be
illustrated by the behaviour of mixtures containing Compimide® 796 and TM 123
BMI resins with GE Ultem 1000®, poly(ether-imide) 26 (Fig. 13) [70]. The critical
stress intensity factor K1C of the linear polymer is six times higher than that of the
BMI matrix and does follow the mixture law for all BMI/Ultem combinations. The
linear polyimide can also be added as 20–40 μm spherical particles to the BMI resin
before it is polymerised. In another example, particles of a soluble precyclised poly-
imide (Ciba-Geigy XU 218®) are dispersed in BMI monomer, which is subsequently
melted at 177°C and cured at 232°C [71]. The two-phase material obtained by this
means exhibits a glass transition temperature of 348°C compared to only 260°C for
the BMI resin without polyimide additive.

The fracture energy values published in the literature have been generally deter-
mined with either neat resins or composite materials and exceptionally with ad-
hesives. The actual fracture energies of thermoplastic and thermosetting adhesives
are, in fact, considerably lower than those measured with moulded resins [72]. This
discrepancy can be explained by the pressure–temperature conditions used in the
bonding techniques. They would not allow the melted resins to flow and wet the
metal surfaces. Moreover, the small thickness of the adhesive joint limits the level
of plastic deformation in contact to the propagating cracks. Table 5 summarises, by
order of increasing energy, the fracture energy values measured at room tempera-
ture for some commercial adhesives.

Table 5. Fracture energies GlC of commercially available high performance adhesives,


measured at ambient temperature [72]

Adhesive trade mark Type of material Glc ( J m− 2)


Narmco SR 5208 Modified epoxy resin 82
Hexcel HX 976 Condensation polyimide 94
Cyanamid FM-300 K Modified epoxy resin 191
Upjohn 2080 Thermoplastic copolyimide 310
Cyanamid FM 34B-18 Condensation oligoimide 385
Larc 13/AATR CTBN-modified bisnadimide 387
Torlon 4000 T Poly(amide-imide) 480
Du Pont NR 056X 6FDA-based polyamic acid 620
Hughes HR 602 Ethynyl-terminated oligoimide 815
Plastilock 650 Nitrile rubber modified pheno- 1037
lic
Plastilock 655 Nitrile rubber modified pheno- 1513
lic
Udel PI 700 Thermoplastic polysulfone 1620
Cyanamid FM 73 Modified epoxy resin 2107

Melt processable thermoplastic adhesives, such as Udel P1700® (G1C = 306 J m− 2),


Upjohn 2080® and Torlon 4000 T® have fracture energies similar to that of the
modified nadimide-terminated polyimide Larc 13/AATR® containing 20% of acry-
lonitrile-butadiene rubber. Compared to the melted form, adhesive films of Udel
P1700® exhibit a fracture energy close to 1.7 kJ m− 2. In the group of thermosetting
adhesives, Cyanamid FM 73® and Plastilock 650–655® possess the best properties of
fracture resistance, surprisingly followed by the ethynyl-terminated HR 602. These
data show that judiciously formulated thermosetting polymers can provide excellent
adhesives.

Commercial bismaleimide adhesives are generally blends of BMI resin, various


nucleophilic co-reactants, and toughening compounds. A number of chemical sys-
tems were studied and developed by Rhône-Poulenc, General Electric, American
Cyanamid and many other companies but the thermal stability of BMI- based
adhesives does not exceed 230°C for long-term uses. Fig. 26 illustrates some data
published by American Cyanamid for A1/A1 assemblies bonded with FM 32 BMI
adhesive cured at 177°C for 4 h under 0.28 MPa [73]. The two other curves refer to
BMIs from Dexter-Hysol Corporation [74]. Aluminium coupons were bonded with
EA 9655 under 0.17 MPa at 177°C for 1 h, followed by 2 h at 232°C, whereas the
bonding conditions for LR 100–774 were 0.2 MPa, 1 h at 177°C and 2 h at 246°C. The
resistance to humidity for A1/FM 32/A1 structures was tested after 72 h in boiling
water, providing lap shear strengths of 20.3, 20 and 19 MPa at 25, 177 and 203°C,
respectively.

Figure 26. Lap-shear strength versus temperature measured on Al/Al assemblies


bonded with: (a) American Cyanamid FM 32 [73], (b) Dexter-Hysol EA 9655 and (c)
Dexter-Hysol LR100-774 [74] BMI adhesives.

> Read full chapter

Maleimide-Based Alder-Enes
M. Satheesh Chandran, C.P. Reghunadhan Nair, in Handbook of Thermoset Plastics
(Third Edition), 2014

Thermoplastic-Toughened Alder-Ene Polymers


Generally, engineering thermoplastics, if blended with bismaleimides, are expect-
ed to impart great improvement in toughness. The drawbacks of thermoplastic
modification include: (i) induced viscosity increase by blending, which hampers
processability of the blend; and (ii) low Tg associated with thermoplastics, which
would be unfavorable for the high-temperature resistance of the blend. There have
been some approaches for the thermoplastic modification of Alder-ene polymers by
various groups.
Diallyl bisphenol A–formaldehyde co-polymer (ABPF, Scheme 12.30) was addi-
tion-cured with bisphenol A–bismaleimide (BMIP) via the Alder-ene reaction by
Nair et al. [106]; its adhesive characteristics were studied by determining lap shear
strength (LSS) and T peel strength.

Scheme 12.30. Diallyl bisphenol A–formaldehyde co-polymer [106].

One-hundred percent retention of the LSS at 250°C was observed for the system
at an optimized 1:1 ratio of allyl to maleimide. The same matrix, when modified
with thermoplastic toughening agents such as polysulfone (PS) and polycarbonate
(PC), reduced the Tg from 350–380°C to 190–260°C. Among these thermoplastic
modifiers (compared to PS), the PC-modified system was less stable at lower temper-
atures, but showed better performance at temperatures above 600°C. The adhesive
characteristics of the different combinations are shown in Table 12.18.

Table 12.18. Thermoplastic-Modified (PS and PC) Adducts of BMIP:ABPF (1:1) [106]

LSS (MPa) Relative Increase Relative Reten-


in LSS (%) tion of LSS (%)
Sys- RT 150°- 200°- 250°- RT 150°- 200°- 250°- 150°- 200°- 250°- TPS
tem C C C C C C C C C (k
Nm−1-
)
Un- 4.1±0.64.- 5.- 5.- – – – – 117 122 127 poor
modi-
fied 8±0.- 0±0.- 2±0.-
4 4 3
PS-10 13.9±0.7
11.- 7.- 4.- 239 129 42 −12 79 51 33 0.-
0±0.- 1±0.- 6±0.- 32±0.-
6 6 5 05
PS-20 19.3±0.7
16.- 11.- 6.- 371 244 120 23 86 57 33 0.-
5±0.- 0±0.- 4±0.- 38±0.-
8 6 5 04
PS-30 14.8±0.6 261 185 114 6 93 72 37
13.- 10.- 5.- 0.-
7±0.- 7±0.- 5±0.- 40±0.-
7 6 6 04
PC-10 11.3±0.8
9.- 8.- 8.- 176 98 68 54 84 74 71 poor
5±0.- 4±0.- 0±0.-
7 7 5
PC-20 8.8±0.86.- 6.- 5.- 115 44 24 4 78 70 61 poor
9±0.- 2±0.- 4±0.-
7 6 6
PC-30 8.3±0.76.- 5.- 3.- 102 40 10 −33 81 66 42 poor
7±0.- 5±0.- 5±0.-
6 6 5

Cured at 160°C for 30 Min, 200°C for 30 Min, and 250°C for 2 H.

A three-component Alder-ene resin composed of BMPM, DABA, and o,o’-dimethal-


lyl bisphenol A (1.0/0.3/0.7 eq ratio) was modified with polyether-ether-ketone
(PEEK) by Takao Iijima et al. [107]. Considerable improvement in toughness was
achieved by the incorporation of PEEK as it provided a co-continuous phase structure
from a single phase structure, as shown in Figure 12.15. Thus, when 15 wt.% of PEEK
was incorporated (44 mol.% TP, MW 23,400), the KIC value for the modified resin
increased to 30% with no penalty in the mechanical and thermal properties. Also,
the KIC increased to 95% when compared to the value for a commercial bismaleimide
(Matrimid 5292™) resin. In a similar work, a BMPM/DABA system was incorporated
with amorphous thermoplastic bisphenol A polysulfone (PSF), polyether ketone
(PEK-C), and polyether sulfone (PES-C) bearing a phthalidylidene group [108]. The
absence of endothermic peak, or the decrease in the area of the endothermic peak
of thermoplastic components in the aged blends, was attributed to the formation of
semi-interpenetrating polymer networks that restricted the segmental mobility of
thermoplastic components.

Figure 12.15. Scanning electron micrographs of the fracture surfaces for the (a) neat,
and (b) 15 wt.% PEEK modified BMPM/DABA system [107].
(Reprinted with permission from Journal of Applied Polymer Science, Vol. 82,
2991–3000, (2001), John Wiley and Sons Publishers.)

In another study, a BMPM/DABA system was modified with 4,4’-bismaleimide


diphenyl ether of biphenyl A (MEBMI), allyl phenol epoxy (APE), and thermoplas-
tic-modified polyether-ketone PEK-C [109]. The authors found that the impact
strength of the BMI resin depended strongly on the amount of PEK-C incorporated.
The typical MBMI/DABPA component system modified with MEBMI, AE, and PEK-C
exhibited outstanding impact toughness, excellent thermal stability, relatively high
Tg, and good processability.

Polyetherimide (PEI) synthesized from different diamines was also used to toughen
BMI [88]. All the modified compositions exhibited two relaxation peaks or shoulders
corresponding to the relaxation transition of the two-blend components, whereas
the unmodified BMI resin gave only one tan peak at 293°C. The incorporation
of the second component induced phase separation in the system. The GIC values
of some of the modified systems were nearly three times higher than the neat
resin. The change in the light-scattering profile with curing time indicated that the
phase separation mechanism depended on the modifier concentration [110]. Thus,
phase separation took place via the spinodal decomposition mechanism in the PEI
15-phr- and 20-phr-modified system and the fracture energy (GIC) increased with
PEI content in the modified system. In the PEI 15-phr-modified system, the GIC
value was three times greater than that of the unmodified BMI resin. Toughening of
commercial BMI resin with varying proportions of poly(phthalazinone ether ketone)
(PPEK) was also reported [111]. The activation energy of the reaction indicated
that the reaction mechanism remained the same even after the incorporation of
PPEK. The morphology of the cured resin changed from a dispersed structure to a
phase-inverted structure with the increase of PPEK content. Compared to the neat
resin, the fracture toughness of the modified resin exhibited a moderate increase
when PPEK was incorporated.

Maleimide functional novolac phenolic resin (PMF), self-cured and co-cured with a
novolac epoxy resin, was modified as a function of the varying concentrations of
the additives, ranging from 10 to 30 phr of the base resin by three thermoplastic
elastomers: (1) two grades of carboxyl terminated butadiene acrylonitrile copolymer
(CTBN) of different molecular weights; (2) a low-molecular-weight, epoxidized hy-
droxyl-terminated polybutadiene [EHTPB]; and (3) a high-molecular-weight acrylate
terpolymer containing pendant epoxy functionality [EPOBAN] (Scheme 12.31). The
adhesive characteristics were studied by determining T peel strength and LSS on
aluminum adherends by Nair et al. [112]. All the elastomers were effective in increas-
ing the adhesive properties at ambient temperature of the brittle, less cross-linked,
self-cured PMF. CTBN-S (high-molecular-weight Mn-65000g/mol) was the most
effective additive with a good phase-separated morphology. For the more rigid, less
ductile, epoxy-cured PMF system, the adhesive properties were marginally improved
by the high-molecular-weight CTBN, whereas the other elastomers were practically
ineffective.

Scheme 12.31. Structures of elastomeric additives [112].

> Read full chapter

Polymers for aerospace structures


In Introduction to Aerospace Materials, 2012

13.5.3 Polyimide, bismaleimide and cyanate


Thermosetting polymers that are also used in structural fibre composites are poly-
imides, bismaleimides and cyanate esters. These polymers are used in aircraft
composite structures required to operate at temperatures above the performance
limit of epoxy resin, which is usually in the upper range of 160–180 °C. Polyimides
can operate continuously at temperatures up to 175 °C and have an operating limit
of about 300 °C. The polyimide called PMR-15 is the most common, and is used
as the matrix phase of carbon– fibre composites in high-speed military aircraft and
jet engine components. The down-side of using polyimides is their high cost. Bis-
maleimide (BMI) is also used in fibre composites required to operate at temperature,
with an upper service temperature of about 180 °C. Carbon–BMI composites are
used in the F-35 Lightning II fighter along with carbon–epoxy materials. Cyanate
resins, which are also known as cyanate esters, cyanic esters or triazine resins, have
good strength and toughness at high temperature, and their maximum operating
temperature is approximately 200 °C. However, cyanate resins pose a safety risk
because they produce poisonous hydrogen cyanide during the cure reaction process.
> Read full chapter

High-Performance Polyimides and


High Temperature Resistant Polymers
Kreisler S.Y. Lau, in Handbook of Thermoset Plastics (Third Edition), 2014

Maleimide-Terminated Thermosetting Polyimides


Another group of thermosetting polyimides curable by addition reactions is the
N-substituted bismaleimides (BMIs). By choosing the appropriate moieties in the
polymer chain, it is possible to prepare very soluble and fusible oligomers. These
BMI prepolymers can be heated above their melting point in the presence of a
free radical catalyst, such as dicumyl peroxide, to polymerize into a crosslinked
polyimide resin. Rhône-Poulenc Company introduced the Kerimid® resins which
are based on mixtures of bismaleimide monomers and aromatic diamines [56].
This results in linear chain extension via a Michael addition reaction. Cross-linked
poly(bismaleimides) were formed from these bismaleimide oligomers. Improved
high-temperature properties could be obtained from these resins if nonstoichio-
metric mixtures of aromatic diamines and bismaleimide were used. Two types
of reactions are postulated [57]. One type is the Michael addition to form linear
polymers as depicted in Figure 10.31. The second type is free radical cure of the
terminal double bonds leading to cross-linking.

Figure 10.31. Chain extension of bismaleimides via Michael addition.

A melt processible bismaleimide copolymer (Kerimid 353) was developed using a


ternary mixture of aliphatic and aromatic bismaleimides [58–60]. Further develop-
ments of bismaleimide technology using a combination of free radical cure and
diamine addition had yielded new types of processible bismaleimide resins, the
Kinels and Kerimid® 601 series, which are suitable as molding and laminating resins,
respectively [61]. Copolymerization studies have demonstrated that bismaleimide
monomers or prepolymers can best be used as cross-linking agents to yield products
with high glass transition temperatures (Figure 10.32) [62].
Figure 10.32. Application of bismaleimides as cross-linking agent.

Table 10.42 shows typical molding parameters of these resins. Thus Kerimid®
FE70003 coated on graphite or glass cloth fibers can be molded using vacuum bag
techniques. Properties of the neat resin are shown in Table 10.43. The low viscosity
and reasonable gel times are very important parameters that make these resins easy
to process. Tables 10.44, 10.45, and 10.46 show typical thermal mechanical and
electrical properties for Kerimid® 601/181E glass cloth laminates.

Table 10.42. Kerimid® FE 70003 Modified Bismaleimides Resin [63]

(Reprinted by permission from Rhône-Poulenc, Inc.)

Table 10.43. Kerimid® FE 70003 Modified Bismaleimide Resin Select Properties of


Neat Resin [63]

Solubility: Solutions of 50% in weight in methylene chloride or methylethyl ketone.

Table 10.44. Kermid® 601/181E Glass Cloth Laminate—Electrical Properties as aRe-


sult of Aging [64]

(Reprinted by permission from Rhône-Poulenc Inc.)

Table 10.45. Kerimid® 601/181E Glass Cloth 18-Ply Laminate—Thermal Aging and
Mechanical Properties [64]

The specimens were taken from laminates prepared in the following conditions:-
Impregnation bath: Solution of Kermid® 601 at 45% in NMP; Fiberglass fabric:
Continuous filament yarn, satin weave of the 181 type with an aminosilane finish;
Resin content in the prepreg: 30 to 35%; Rate of flow of the prepreg: 30 to 40%;
Laminate thickness: Stack of 18 plies; Curing conditions: (a) Under 210 psi pressure
at 480°F and postcuring for 48 hours at 390°F, or (b) Curing under 210 psi at 390°F
and postcuring for 24 hours at 480°F; Resin content: 22 to 24%; Specific gravity:
1.94; Barcol Hardness: ca 70.

(Reprinted by permission from Rhône-Poulenc Inc.)

Table 10.46. Selected Properties of Kerimid® 601/181E Glass Cloth Laminates [64]

(Reprinted by permission from Rhône-Poulenc Inc.)

Since many years ago, Ciba-Geigy marketed a two-component bismaleimide system


XU292 which, when combined and cured, is suitable for high temperature advanced
composites and adhesives applications. The system is based on 4,4 -bis(maleimi-
dophenyl)methane (III) and 3,3 -diallyl Bisphenol A (XXII)[65] (Figure 10.33).

Figure 10.33. Components of Ciba-Geigy XU292 BMI resin (Matrimid® 5292).

Compounds III and XXII are well known as the Matrimid® 5292 system (III=5292A
and XXII=5292B), complementing the polyimide Matrimid® 5218. Upon combining
the two components with continuous stirring to 120°–150°C (248°–302°F), a clear
homogeneous solution is obtained. The resulting liquid can be used either as a
casting resin for the preparation of prepregs or as an adhesive. The two-component
system provides flexibility to provide the optimum formulation for the prepregger.
One formulation of the neat resin reported by Ciba-Geigy [66] reports a room
temperature tensile strength of 13.6 ksi, a modulus of 564 ksi, an elongation of
3.0%, a flexural strength of 26.8 ksi and a flexural modulus of 580 ksi. At 204°C
(400°F), these values drop to 104 ksi for tensile strength and 394 ksi for the modulus.
The Tg obtained from the TMA penetration method is 282°C (540°F).

In many oil field downhole operations, service tools are required to perform in a
hot/wet fluid environment with temperatures above 204°C (400°F) and pressures
above 70 MPa (ca. 10 KSI) [67]. It appears that bismaleimides are well suited for this
application.

Other similar bismaleimide (BMI) systems have also been developed as Ciba’s BMI
family of high-performance resins. A notable new BMI is RD85-101, which was syn-
thesized from 1,1,3-trimethyl-3-phenylindane and maleic anhydride (Figure 10.34).
Formulated with common BMI co-curing agents such as allyl phenols (e.g., Matrim-
id® 5292B) or aromatic diamines, RD85-101 shows outstanding thermomechanical
and hygrothermal peorformance. Table 10.47 shows some exemplary physical and
thermomechanical data of this resin system [68].

Figure 10.34. RD85-101, BMI Based on 1,1,3-Trimethyl-3-phenylindane.

Table 10.47. Cured Neat Resin Thermomechanical Data of RD85-101/Matrimid


5292B Formulations [68]

Cure condition: 1 hr/180°C+2 hr/200°C+6 hr/250°C.Immersion in 160°F water for 48


hours.System I=1:1 molar ratio (RD85-101/Matrimid 5292B)System II=1:0.87 molar
ratio (RD85-101/Matrimid 5292B)

> Read full chapter

High-Performance Polyimides and Re-


lated Thermoset Polymers: Past and
Present Development, and Future Re-
search Directions
Abraham L. Landis, Kreisler S.Y. Lau, in Handbook of Thermoset Plastics (Second
Edition), 1998

MALEIMIDE-TERMINATED THERMOSETTING POLYIMIDES


Another group of thermosetting polyimides curable by addition reactions is the
N-substituted bismaleimides (BMIs). By choosing the appropriate moieties in the
polymer chain, it is possible to prepare very soluble and fusible oligomers. These
BMI prepolymers can be heated above their melting point in the presence of a
free radical catalyst, such as dicumyl peroxide, to polymerize into a cross-linked
polyimide resin. Rhône-Poulenc Company introduced the Kerimid® resins which are
based on mixtures of bismaleimide monomers and aromatic diamines.[49] Such mix-
tures produce linear chain extension via a Michael addition reaction. Cross-linked
poly(bismaleimides) were formed from these bismaleimide oligomers. Improved
high-temperature properties could be obtained from these resins if nonstoichio-
metric mixtures of aromatic diamines and bismaleimide were used. Two types of
reactions are postulated.[50] One type is the Michael addition to form linear polymers
as depicted in Figure 8.31. The second type is free radical cure of the terminal double
bonds leading to cross-linking.

Figure 8.31. Chain extension of bismaleimides via Michael addition.

A melt processible bismaleimide copolymer (Kerimid 353) was developed using a


ternary mixture of aliphatic and aromatic bismaleimides.[51]–[53] Further develop-
ments of bismaleimide technology using a combination of free radical cure and
diamine addition had yielded new types of processible bismaleimide resins, the
Kinels and Kerirnid® 601 series, which are suitable as molding and laminating
resins, respectively.[54] Recent copolymerization studies have demonstrated that
bismaleimide monomers or prepolymers can best be used as cross-linking agents
to yield products with high glass transition temperatures (Figure 8.32).[55]

Figure 8.32. Application of bismaleimides as cross-linking agent.

Typical molding parameters of these resins are presented in Table 8.42. Thus,
Kerimid® FE70003 coated on graphite or glass cloth fibers can be molded using
vacuum bag techniques. Properties of the neat resin are listed in Table 8.43. The low
viscosity and reasonable gel times are very important parameters which make these
resins easy to process. Tables 8.44 through 8.46 show typical thermal mechanical
and electrical properties for Kerimid® 601/181E glass cloth laminates.

Table 8.42. Kerimid® FE 70003 Modified Bismaleimides Resin [56]

Processing
Impregnation Reinforced from glass cloth, graphite, etc., can be
coated from melted resin (90°C) or from a lacquer
in methylene chloride or methylethyl ketone.
Molding Can be molded by vacuum bag technique. The
cure takes place at 200° to 250°C at pressures of
75 to 150 psi. It is advisable to postcure the parts
12 to 24 hours at 250°C
Properties of Composites Seven-plie graphite fabric at 215g/m2 with 42%
resin content and molded and cured as indicated
above
Flexural strength, 103 × psi 98.6 (20°C, 68°F)
Flexural modulus, 106 × psi 0.77 (20°C, 68°F)

(Reprinted by permission of Rhône-Poulenc, Inc.)

Table 8.43. Kerimid® FE 70003 Modified Bismaleimide Resin Select Properties of


Neat Resin [56]

Property Value
Density, g/cc 1.2
Viscosity, cps
60°C 27,000
70°C 4,500
80°C 1,200
90°C 500
100°C 450
Gel Time, min
150°C 45
180°C 20

Note:Solubility: Solutions of 50% in weight in methylene chloride or methylethyl


ketone.

Table 8.44. Electrical Properties of Kermid® 601/181E Glass Cloth Laminate as a


Result of Aging [57]

Condition Dielectric Volume Resistivi- Dielectric Con- Dissipation Factor,


Strength, kV/mm ty, ohm × cm stant, 1 kHz 1 kHz
ASTM method D149 D257 D159 D150
Initial 25 6×1014 4.5 0.012
24 hr. in water 20 1.5×1013 5.4 0.016
1000 hr. at 355°F &gt;16.5 — — —
1000hr. at 390°F &gt;16.5 — — —
1000 hr. at 430°F 12 — — —
2000 hr. at 480°F — 2.2×1015 — —
10000 hr. at 355°F — — 5.5 —
10000 hr. at 390°F — — 5.5 —
10000 hrs at 430°F — — 4.7 —

(Reprinted by permission of Rhône-Poulenc Inc.)

Table 8.45. Thermal Aging and Mechanical Properties of Kerimid® 601/181E Glass
Cloth 18–ply Laminate [57]

Flexural Strength, Flexural Modulus,


psi × 103 psi × 103
Temperature 77°F 390°F 77°F 390°F Weight Loss,
of Aging, °F, %
hr.
180 Initial 71 57 3850 3150 —
1000 67 59 3600 3350 0.25
2000 67 59 3800 3350 0.30
5000 67 57 3850 3400 0.5
8000 59 52 3500 3300 0.9
10000 59 47 3850 3350 1.4
390 Initial 71 57 3850 3150 —
1000 66 56 3650 3400 0.4
2000 66 54 3650 3400 0.5
5000 57 54 3650 3350 1.1
8000 46 37 3300 3150 1.9
10000 37 30 3600 3000 2.7
430 Initial 71 57 3850 3150 —
1000 64 56 3800 3400 0.7
2000 64 54 3700 3200 0.9
3000 59 51 3600 3200 1.4
5000 44 37 3300 3050 2.8
8000 19 16 2800 2600 5.4
10000 17 12 1800 2450 7.7
480 Initial 71 57 3850 3150 —
1000 57 47 3300 3150 4.0
2000 50 44 3100 2950 5.7
3000 26 23 2600 2600 8.3

The specimens were taken from laminates prepared in the following conditions:Im-
pregnation bath:Solution of Kermid® 601 at 45% in NMPFiberglass fabric:Continuous filament yarn,
satin weave of the 181 type with an aminosilane finishResin content in the prepreg:30 to 35%Rate of
flow of the prepreg:30 to 40%Laminate thickness:Stack of 18 plies;Curing conditions:(a) Under 210
psi pressure at 480°F and postcuring for 48 hours at 390°F, or(b) Curing under 210 psi at 390°F and
postcuring for 24 hours at 480°F.Resin content:22 to 24%Specific gravity:1.94Barcol Hardness:370.
(Reprinted by permission of Rhône–Poulenc Inc.)

Table 8.46. Selected Properties of Kerimid® 601/181E Glass Cloth Laminates [57]

Property ASTM Approximate Value


Flexural strength, psi × 103
77°F (25°C) 70
390°F (200°C) 60
480°F (250°C) 50
Flexural modulus, psi × 103 D790
77°F (25°C) 4000
390°F (200°C) 3800
480°F (250°C) 3200
Tensile strength, psi × 103 D638
77°F (25°C) 50
Compressive strength, psi × 103 D695
77°F (25°C) 50
Delaminating strength, psi
77°F (25°C) D2355 2150
Izod Impact Strength, ft × lb/in D256
77°F (25°C)
Notched 13
Unnotched 15

(Reprinted by permission of Rhône-Poulenc Inc.)

Several years ago Ciba-Geigy marketed a two-component bismaleimide system


XU292 that, when combined and cured, is suitable for high temperature advanced
composites and adhesives applications. The system is based on 4,4 -bis(maleimi-
dophenyl)methane (III) and 3,3 -diallyl Bisphenol A (XXII)[58] (Figure 8.33).

Figure 8.33. Components of Ciba–Geigy XU292 BMI resin (Matrimid® 5292).

Upon combining the two components with continuous stirring to 120° to 150°C
(248° to 302°F), a clear homogeneous solution is obtained. The resulting liquid can
be used either as a casting resin for the preparation of prepregs or as an adhesive.
The two-component system provides flexibility to ensure the optimum formulation
for the prepreg. One formulation of the neat resin by Ciba-Geigy[59] reports a room
temperature tensile strength of 13.6 ksi, a modulus of 564 ksi, an elongation of
3.0%, a flexural strength of 26.8 ksi, and a flexural modulus of 580 ksi. At 204°C
(400°F), these values drop to 104 ksi for tensile strength and 394 ksi for the modulus.
The Tg obtained from the TMA penetration method is 282°C (540°F).

Compounds III and XXII, now known as the Matrimid® 5292 system (III=5292A
and XXII=5292B), complements the polyimide Matrimid® 5218. Other similar bis-
maleimide (BMI) systems have also been developed as Ciba's BMI family of high-per-
formance resins. A notable new BMI is RD85-101, that was synthesized from
1,1,3-trimethyl-3-phenylindane and maleic anhydride (Figure 8.34). Formulated
with common BMI co-curing agents such as allyl phenols (e.g., Matrimid® 5292B)
or aromatic diamines, RD85-101 shows outstanding thermomechanical and hy-
grothermal performance. Table 8.47 shows some exemplary physical and thermo-
mechanical data of this resin system.[60]
Figure 8.34. RD85–101, BMI based on 1,1,3–trimethyl–3–phenylindane.

Table 8.47. Cured Neat Resin Thermomechanical Data of RD85-101/Matrimid


5292B Formulations [60]

Test Temperature
Property 25°C 177°C 232°C
Flexural strength, ksi
System I 17.5 15.9 13.2
System II 17.5 16.2 (204°C)
Flexural modulus, ksi
System I 531 428 348
System II 502 383 (204°C)
Tensile strength, ksi
System I 11.5 — —
System II 8.8 7.8 (204°C) —
Tensile modulus, ksi
System I 539 — —
System II 547 396 (204°C)
Tensile elongation, %
System I 1.8 — —
System II 1.8 2.0 (204°C)
DMA modulus (dry), ksi 512 402 334
(System II) (150°C) (250°C)
DMA modulus
(hot/wet)*, ksi
(System II) 463 336 (150°C) (250°C)
Tg, °C (TMA), System II 298
Water Pick-up*, System 2.6%
II

Note:Cure condition: 1 hr/180°C + 2 hr/200°C + 6 hr/250°CSystem I=1:1 molar ratio


(RD85-101/Matrimid 5292B)System II = 1:0.87 molar ratio (RD85-101/Matrimid
5292B)

* Immersion in 160°F water for 48 hours

> Read full chapter


Morphology and Properties of Polyben-
zoxazine Blends
Chongyin Zhang, ... Sixun Zheng, in Handbook of Benzoxazine Resins, 2011

Publisher Summary
Polybenzoxazines are a class of attractive alternatives to some traditional thermosets
such as epoxies, phenolic resins, and bismaleimides owing to their excellent thermal,
mechanical, and electrical properties. The thermosets can be prepared via thermally
activated ring-opening polymerization of benzoxazine monomers and no hardeners
are required in the polymerization process. The unique chemistry of polymerization
endows the materials with excellent processing properties through a very wide range
of molecular design. Over the past few decades, considerable progress is made in
the synthesis of a variety of precursors to improve the properties of this class of
thermosets. It is recognized that besides the development of new polybenzoxazine
materials through the chemical syntheses of a series f new benzoxazine monomers,
benzoxazine resins can alternatively be modified via a blending approach. The
polybenzoxazines can be modified by incorporating a variety of thermoplastics or
elastomers into the thermosets. The control over the miscibility and morphology
of polybenzoxazine blends is important for the optimization of the intercompo-
nent interactions, thereby endowing the materials with improved properties. It is
identified that the morphological structures of polybenzoxazine blends are quite
dependent on the intermolecular specific interactions and the competitive kinet-
ics between polymerization and phase separation. As a proton-donating polymer,
polybenzoxazine can be miscible with some proton-accepting polymers such as
poly(N-vinylpyrrolidone) because of the formation of the favorable intermolecular
hydrogen bonding interactions. The formation of fine heterogeneous morphology
in polybenzoxazine thermosets has a profound impact on the thermal, mechanical,
optical, and processing properties of the materials. For elastomer-modified polyben-
zoxazine thermosets, the improvement of fracture toughness follows several known
toughening mechanisms such as shear yielding, particle bridging, crack-pinning,
and microcracking mechanisms.

> Read full chapter

Various Approaches for Main-Chain


Type Benzoxazine Polymers
Saeed Alhassan, ... Hatsuo Ishida, in Handbook of Benzoxazine Resins, 2011
1 Introduction
Polybenzoxazines are a class of thermosetting resins that compare favorably with
traditional thermosets, such as phenolic, epoxy, bismaleimides, and cyanate ester
resins, and cross-linkable polyimides with additional unusual properties. The unique
properties of benzoxazines make them ideal for various applications where these
traditional monomers have failed. Among these properties are low water absorp-
tion, extremely rich molecular design flexibility, fast property development at low
conversion, and near-zero shrinkage upon polymerization [1,2]. The early research
on the chemistry of polybenzoxazines focused on the monomeric type, where a
monofunctional amine, phenol, and formaldehyde react in a Mannich condensation
reaction to release water and afford a classical benzoxazine monomer, as shown
in Scheme 1. Upon heating to 160-220 °C, benzoxazine undergoes ring-opening
polymerization, and cross-linked polybenzoxazine is produced. Schemes 2 and 3
show two main classes of difunctional benzoxazine monomers: the combination
of bisphenol and monoamine, or diamine with a monophenolic derivative. Mono-
functional benzoxazines lead to small oligomers upon polymerization and do not
produce structurally strong, cross-linked polymers, unless they are combined with
groups that polymerize in non-benzoxazine chemistry. An extension to these difunc-
tional, monomeric precursors is the use of bisphenols and aromatic diamines, which
lead to formation of a linear polymer with benzoxazine groups in the main chain,
as shown in Scheme 4, leading to the enhancement of the overall properties of the
thermoset. Cross-linked polybenzoxazines derived from monomers are brittle as is
the case for all thermosetting resins. Propagation of polymer chains competes with
intramolecular hydrogen bond formation, leading to the tendency to terminate and
branch the network structure. These branches were hypothesized to be the cause
of the low temperature degradation. To minimize these problems, the main chain
benzoxazine polymer (MCBP) approach, which leads to more ductile cross-linked
polybenzoxazines and enhanced thermal properties, was developed.

Scheme 1. Classical benzoxazine monomer of aniline and phenol (P-a) and the
resulting polybenzoxazine.

Scheme 2. Synthesis of bisphenol-based benzoxazine monomer.


Scheme 3. Synthesis of diamine-based benzoxazine monomer.

Scheme 4. Structure of MCBP [4,5].

MCBPs have been developed as a new generation of polybenzoxazine precursors.


Liu et al. [3] have conducted a study to synthesize MCBP using 4,4 -methyl-
enebis(2,6-dimethylaniline), bisphenol-A, and formaldehyde. Takeichi et al. [4] and
Chernykh et al. [5] published detailed studies of the synthesis of high molecular
weight MCBP through the polycondensation of diamines and bisphenols to produce
an AA BB-type linear polymer with benzoxazine rings in the main chain (Scheme 4),
hereinafter called main-chain type benzoxazine polymers (MCBPs). Many MCBPs
have been developed over the following years. Benzoxazine monomers have small
molecular weights and are thus difficult to process into self-supporting films.
Due to their higher molecular weight, the MCBPs can be easily processed as ther-
moplastics in melt processing equipment. In this chapter, the different strategies
of synthesizing MCBP are discussed along with key properties of the synthesized
MCBP.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

Vous aimerez peut-être aussi