Вы находитесь на странице: 1из 17

HIGHLIGHT

Dendritic Aromatic Polyamides and Polyimides


MITSUTOSHI JIKEI,1 MASA-AKI KAKIMOTO2 1 Department of Material-Process Engineering and Applied Chemistry for Environments, Akita University, Tegata Gakuen-Machi, Akita-Shi, Akita 010-8502, Japan Department of Organic and Polymeric Materials, Tokyo Institute of Technology, MeguroKu, Tokyo 152-8550, Japan Received 5 December 2003; accepted 8 December 2003
2

ABSTRACT: The syntheses and properties of dendritic and hyperbranched aromatic polyamides and polyimides are reviewed. In addition to conventional stepwise reactions for dendrimer synthesis, an orthogonal/double-stage convergent approach and dendrimer syntheses with unprotected building blocks are described as new synthetic strategies for dendritic poly-

amides. Hyperbranched polyamides and polyimides composed of various repeating units are presented. Besides the self-polycondensation of AB2-type monomers, new polymerization systems with AB4, AB8, A2 B3, and A2 BB2 monomers have been developed for hyperbranched polyamides and polyimides. The copolymerization of AB2 and AB

monomers is discussed separately with respect to the effects of branching units on the properties.
2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 12931309, 2004

Keywords: dendrimers; hyperbranched; aromatic polyamides; polyimides; synthesis; copolymerization

DR. MITSUTOSHI JIKEI

Dr. Mitsutoshi Jikei received his B.S. in 1988, his M.S. in 1990, and his Ph.D. in 1993 from Waseda University under the guidance of Professor E. Tsuchida on the subject of the oxidative polymerization of aromatic thioether polymers. In 1993, he joined the Department of Organic and Polymeric Materials at the Tokyo Institute of Technology and started his academic carrier with Professor M. Kakimoto. He worked as a postdoctoral fellow for Professor Joseph M. DeSimone at the University of North Carolina from 1995 to 1996. In 2003, he moved to Akita University and worked as an associate professor in the Department of Material-Process Engineering and Applied Chemistry for Environments. Since 1994, he has pursued research on dendritic and hyperbranched aromatic polymers, such as dendritic polyamides, hyperbranched polyamides, polyimides, and polyesters. He has published about 90 original publications and review articles. He has received an award for the outstanding paper published in the Polymer Journal in 2002. His current research interests include the design, preparation, and properties of functional organic materials.

Correspondence to: M.-A. Kakimoto (E-mail: mkakimot@o. cc.titech.ac.jp)


Journal of Polymer Science: Part A: Polymer Chemistry, Vol. 42, 12931309 (2004) 2004 Wiley Periodicals, Inc.

1293

1294

J. POLYM. SCI. PART A: POLYM. CHEM.: VOL. 42 (2004)

DR. MASA-AKI KAKIMOTO

Dr. Masa-Aki Kakimoto received his Ph.D. in 1980 from the Tokyo Institute of Technology on the subject of sigmatropic rearrangement. He joined Professor Larry E. Overmans group at the University of California at Irvine from 1977 to 1978 as a visiting graduate student. After he received his Ph.D., he spent 2 years at the Sagami Central Chemical Research Center, studying the synthesis of antibiotics. He moved to the Department of Organic and Polymeric Materials at the Tokyo Institute of Technology, starting his academic carrier with Professor Yoshio Imai in 1982. He was promoted to an associate professorship in 1987 and to a full professorship in 1997. Since he moved to the Tokyo Institute of Technology, he has studied condensation polymers, such as new high-temperature polymers, LangmuirBlodgett lms of polyimides, and polyimide silica composites via the sol gel process. Currently, he is focusing on dendritic polymers. He has published about 480 original publications and review articles. He received an Ichimura award in 1996 and an award from the Society of Fiber Science and Technology in 2000.

INTRODUCTION
Dendritic macromolecules are classied as dendrons, dendrimers, and hyperbranched polymers, which are composed of successive branching units. Unique properties, such as good solubility, low viscosity, multivalence, and encapsulation effects, have been reported for many dendritic macromolecules, and they are mainly caused by the branching and spherical architecture. In addition to dendrimers composed of exible chains, such as poly(amidoamine), poly(propylene imine), aliphatic polyester dendrimers, and hyperbranched polyesters, dendritic branching architecture has been incorporated into rigid aromatic polymers. Dendritic aromatic polyethers, polyesters, polyamides, polyimides, polysuldes, and polysulfones have already been reported in the literature.17 Linear aromatic polyamides (aramids) and polyimides are known as important high-performance polymers because of their excellent thermal, mechanical, and chemical properties. Aramid bers, prepared by liquid-crystal spinning, and lms have a large number of applications in modern industries. Linear aromatic polyimides are among the most thermally stable polymers to date, and they are useful for specialty applications in, for example, the microelectronics and aerospace industries. Because of rigid repeating units, aramids and linear polyimides are practically insoluble in organic solvents at room temperature. The introduction of the dendritic structure might produce new processable high-performance materials. This highlight deals with dendritic aromatic polyamides and polyimides mainly from a synthetic point of view. Recent progress in the synthesis of dendritic and hyperbranched polyamides and polyimides is reviewed

in addition to conventional stepwise reactions and selfpolymerization. The copolymerization of AB2 and AB monomers is discussed separately to discuss the effect of branching units on the properties.

PREPARATION OF DENDRITIC POLYAMIDES


Polyamide dendrimers were rst reported by Miller and Neenan in 1990.8 The synthetic pathway was based on a convergent approach. 1,3,5-Benzenetricarbonyl trichloride and 5-nitroisophthaloyl chloride were used as the core and building blocks, respectively, as shown in Scheme 1. Only brief comments on NMR, mass spectrometry, and gel permeation chromatography (GPC) data were described in the text for the polyamide dendrimers (1 and 2). Feast and coworkers9,10 reported the synthesis and detailed characterization of aromatic polyamide dendrimers (1, 2, and 3). The peak splitting of aromatic protons derived from inner aromatic rings was observed in an 1H NMR spectrum for compound 2, suggesting sterically restricted rotation of inner amide bonds. An attempt to prepare compound 3 in the same manner as 2 produced a poor yield, probably because of steric hindrance. Although compound 1 was highly crystalline, compound 2 was amorphous and showed good solubility in tetrahydrofuran (THF; 298 g/L). We have reported the synthesis and properties of polyamide dendrons and dendrimers by orthogonal and double-stage convergent approachs.11 As shown in Scheme 2, dendrons with an aromatic bromide group (4 and 6) were prepared by the orthogonal approach. 3,5Diaminobromobenzene and 3,5-diaminobenzoic acid were used as building blocks. The direct condensation of

HIGHLIGHT

1295

Scheme 1. Preparation of polyamide dendrimers by divergent and convergent approaches.

carboxylic acid and aromatic amines in the presence of condensation agents and a CO insertion reaction catalyzed by a palladium complex were applied in an alternating fashion. The other dendrons with amino groups on the periphery (7 and 8) were prepared from 3,5-diami-

nobenzoic acid and 3,5-dinitrobenzoyl chloride through a divergent approach (Scheme 3). The application of the double-stage convergent growth approach to 4 and 7 or 8 gave fourth-generation (9) or fth-generation (10) dendrons in a one-step reaction, as shown in Scheme 4.

1296

J. POLYM. SCI. PART A: POLYM. CHEM.: VOL. 42 (2004)

Scheme 2. Orthogonal approach to the preparation of polyamide dendrons.

Because the starting dendrons (4, 7, and 8) could be isolated without column chromatography, these highgeneration dendrons (9 and 10) could be synthesized readily on a large scale. The coupling reaction of the dendrons with a core molecule, such as p-phenylenediamine, gave a corresponding dendrimer (Scheme 5). Similarly to the case reported by Feast and coworkers,9,10 the yield of high-generation dendrimers decreased, probably because of steric hindrance at a focal

point. The structure of the polyamide dendrons and dendrimers was conrmed by 1H NMR, 13C NMR, and matrix-assisted laser desorption/ionization time-of-ight mass spectrometry (MALDI-TOF MS) measurements. The peak assignments of 1H and 13C NMR spectra supported the formation of the proposed structure of the dendrons. A MALDI-TOF MS spectrum of the G5 dendron (10) showed the expected peaks at 7531.7 (M Na) and 7641.5 (M Cs). Although dendron 8,

Scheme 3. Divergent approach to the preparation of amine-terminated dendrons.

HIGHLIGHT

1297

Scheme 4. Double-stage convergent approach to the preparation of polyamide dendrons.

possessing eight functional groups, was used as the building block, the peaks assigned to incompletely substituted products were not detected for G5 dendron 10 after the purication of the crude product by preparative GPC in THF. Recently, Ueda and coworkers1214 reported the synthesis of aromatic polyamide dendrimers from unprotected AB2 building blocks. The key reaction that they applied is the direct condensation of carboxyl and amino groups in the presence of diphenyl(2,3-dihydro2-thioxo-3-benzoxazolyl)phosphonate (DBOP), as shown in Scheme 6. The carboxyl group is converted

into an activated amide by treatment with an equivalent amount of DBOP. The active amide then reacts with an AB2-type molecule to form a branched molecule via a divergent or convergent approach. Selfpolycondensation of the AB2 molecule is not allowed because only the activated amide can react with the amino group of the AB2 molecule. Spacer groups between aromatic rings, such as ether and ethyl groups, ease the steric hindrance of high-generation dendrimers. A dumbbell-shaped dendrimer (12), shown in Figure 1, was prepared by the strategy with an unprotected AB2 monomer via a convergent ap-

Scheme 5. Coupling reaction of polyamide dendrons with p-phenylenediamine.

1298

J. POLYM. SCI. PART A: POLYM. CHEM.: VOL. 42 (2004)

Scheme 6. Direct condensation of carboxyl and amino groups in the presence of DBOP.

proach. The MALDI-TOF MS spectrum of 12 clearly showed the formation of the proposed dendrimer.

PREPARATION OF HYPERBRANCHED POLYAMIDES


There are two major routes to the synthesis of wholly aromatic polyamides: (1) the low-temperature polycondensation reaction of aromatic amines and acid chlorides and (2) the direct polycondensation of aromatic amines and carboxylic acids in the presence of condensation

agents. Kim15 rst reported the synthesis of hyperbranched aromatic polyamides by the low-temperature polycondensation of AB2 and A2B monomers, as shown in Scheme 7. The sulnyl amino acid chlorides prepared from the aromatic amino acids and thionyl chloride were treated with dry hydrochloric acid to yield amino acid chloride hydrochloride. The hydrochloride was neutralized in an amide solvent to form a free amine in situ, and this allowed self-polycondensation to take place. The sulnyl amino acid chloride derivatives were alternatively polymerized through the addition of 1 equiv of water. The GPC analysis of the resulting hyperbranched

Figure 1. Polyamide dendrimer (12) prepared with an unprotected AB2 monomer.

HIGHLIGHT

1299

Scheme 7. Low-temperature polycondensation for the formation of hyperbranched aromatic polyamides.

polyamides in a dimethylacetamide (DMAc)/LiBr/ H3PO4/THF mixture showed that the molecular weight ranged between 24,000 and 46,000 with polydispersities of 2.0 3.2 on the basis of polystyrene standards. Both carboxyl- and amino-terminated polyamides were soluble in amide solvents, such as dimethylformamide (DMF), N-methylpyrrolidone (NMP), and DMAc, whereas the crude amino-terminated polyamide was soluble in water because the amino groups were converted into their hydrochloride salt during the polymerization. The direct polycondensation in the presence of condensation agents for the activation of carboxylic acid groups in situ is the other attractive route to the synthesis of hyperbranched polyamides. A potential advantage of direct polymerization is the stability of monomers at room temperature, which allows many kinds of ABx monomers to be examined as starting materials. Some of the AB2 monomers for hyperbranched aromatic polyamides reported in the literature are listed in Figure 2.16 21 Triphenyl phosphite/pyridine and DBOP/triethylamine have been reported to be efcient condensation agents for polymerization. Polymerization usually proceeds homogeneously in an aprotic polar solvent, similarly to linear aromatic polyamides. All of the hyperbranched polyamides prepared from the AB2 monomers listed in Figure 2 were soluble in amide solvents, and the solubility and thermal properties were dependent on the nature of the end functional groups. In addition, the inherent viscosity of the resulting polymers was low, in the range of 0.1 0.4 dL/g, similar to that of other hyperbranched and dendritic polymers in the literature. The shape factor in the MarkHouwink equation ([] KM) for the hyperbranched polyamide prepared from

13 was 0.35; it was determined from the relationship between the intrinsic viscosity and the molecular weight.17 Hyperbranched polymers contain imperfectly branched linear units in addition to dendritic and terminal ones. Three possible repeating units for the hyperbranched polyamide prepared from 13 are described in Figure 3. The degree of branching (DB) is often used to show the ratio of three units.22,23 The DBs of hyperbranched polymers are generally around 0.5, which is the value anticipated by statistical considerations.23 A quantitative 13C NMR spectrum of the hyperbranched polyamide prepared from 13 gave a DB of 0.46.17 This fact implies that the propagation reaction proceeds statistically without severe steric hindrance and reactivity change during the polymerization. However, the DB of the polymer prepared from 16 was 0.32, which is much lower than the statistical expectation of 0.5.19 The low DB might be caused by steric hindrance during the polymerization. It is clear that the DB of hyperbranched polyamides is greatly inuenced by the structure of the starting monomers. The polymerization of AB4 and AB8 monomers, shown in Figure 4, resulted in the formation of polymers with higher DBs (0.72 for the AB4 monomer and 0.84 for the AB8 monomer).19 The increase in DB is mainly caused by the prefabricated dendritic units and is not directly related to the formation of spherical molecules. DBs could also be controlled by the copolymerization of the AB2 and AB4 monomers continuously from 0.3 to 0.7.19 Although a noticeable inuence from enhanced DBs of the hyperbranched polyamides has not been observed, the DB difference clearly resulted in a

Figure 2. AB2-type monomers for the preparation of hyperbranched polyamides.

1300

J. POLYM. SCI. PART A: POLYM. CHEM.: VOL. 42 (2004)

Figure 3. Three possible repeating units for the hyperbranched polyamides prepared from 13.

different architecture and potentially affected the properties of the polymers, such as the encapsulation effect. According to a statistical study by Flory,24 the polycondensation of difunctional (A2) and trifunctional (B3) compounds results in gelation over a certain conversion of functional groups. Aharoni and coworkers2528 reported the preparation and properties of aromatic polyamide networks. When the polymerization is terminated before a certain conversion, a soluble highly branched polymer can be isolated. Potentially, hyperbranched polymers can be prepared from A2- and B3-type monomers if AB2-type molecules are formed predominantly during the initial stage of the polymerization. Hyperbranched polyamides prepared from A2 and B3 monomers by direct polycondensation have already been reported in the literature.18,29,30 The examined monomers are listed in Figure 5. The resulting hyperbranched polyamides showed good solubility in aprotic polar solvents and multifunctionality, which allowed chemical modication, similarly to hyperbranched polymers prepared from ABx-type monomers. However, the inherent viscosity of the polyamides was high, and this was probably caused by intramolecular crosslinking reactions, intermolecular crosslinking reactions, or both. Komber et al.31 reported a detailed analysis by NMR spectroscopy for the repeating units present in a hyperbranched polyamide prepared from p-phenylenediamine (A2) and tri-

Figure 5. A2 and B3 monomers examined as monomers for hyperbranched polyamides.

Figure 4. AB4 and AB8 monomers for hyperbranched polyamides.

mesic acid (B3). An excess of B3 resulted in a larger number of branches and predominant COOH functionalization. Recently, Chang and Shu32 reported the polymerization of A2 and BB2 monomers for the preparation of hyperbranched aromatic poly(amide imide) (Scheme 8). The rapid reaction between the anhydride and amino group led to the predominant formation of an AB2-type molecule. The direct polycondensation in the presence of condensation agents gave the hyperbranched aromatic poly(amide imide), the structure of which was nearly identical to that prepared from an isolated AB2 monomer. No gelation occurred during the polymerization of the A2 and BB2 monomers. It is generally difcult to obtain aromatic polyamides with a high molecular weight by a molten polycondensation method, although thermal polymerization is a well-known process for aliphatic polyamides with a high molecular weight. However, melt polycondensation is applicable to some AB2-type monomers for the preparation of hyperbranched aromatic polyamides with reasonably high molecular weights, as shown in Scheme 9.33 The thermal polymerization of 3,5-bis(4-aminophenoxy)benzoic acid (13) at 235 C proceeded in a stable molten state and gave a glassy solid after cooling to room temperature. Pouring a DMF solution of the glassy product into methanol containing lithium chloride precipitated a white polymer. The weight-average molecular weight (Mw) and molecular weight distribution [weightaverage molecular weight/number-average molecular

HIGHLIGHT

1301

Scheme 8. A2 BB2 and AB2 polymerization for the formation of the hyperbranched poly(amide imide)s reported by Shu.

weight (Mw/Mn)], determined by gel permeation chromatography/multi-angle laser light scattering (GPC MALLS), were 74,600 and 2.6, respectively. The thermal polymerization of the corresponding AB monomer, 3-(4-aminophenoxy)benzoic acid, failed because the melt phase advanced to a nontransparent solid quickly. The amorphous state and low viscosity due to the branching architecture of the hyperbranched polymer seemed to be suitable for the thermal polymerization. Palladium-catalyzed CO insertion was also attempted as a propagation reaction for the preparation of hyperbranched aromatic polyamides.34 Unfortunately, the polymerization of 3,5-dibromoaniline resulted in the formation of an insoluble product, which might have been caused by undesired side reactions. The polymerization with a core molecule, tetraphenyladamantane, gave a partially soluble product.

PREPARATION OF DENDRITIC AND HYPERBRANCHED POLYIMIDES


There are a few publications on dendritic aromatic polyimides prepared by stepwise reactions. Shu and coworkers35,36 reported the preparation of dendritic poly(ether imide)s by a convergent growth approach, using 1-(4aminophenyl)-1,1-bis(4-hydroxyphenyl)ethane as a building block (Scheme 10). The aromatic nucleophilic substitution of the building block with 3-nitro-N-phe-

nylphthalimide led to the rst-generation dendron (18) with an aminophenyl group at the focal point. The amino group was reacted with 3-nitrophthalic anhydride to yield 19 with an activated nitro group. The secondgeneration dendrimer was prepared by the coupling reaction of the second-generation dendron with 1,1,1tris(4-hydroxyphenyl)ethane. The structure of the dendrons and dendrimers was conrmed by 1H NMR, 13C NMR, IR, and mass spectrometry. In addition, hydroxylterminal, benzyl ether terminated, and hexyl-terminated dendrons and dendrimers were prepared in the same manner. The glass-transition temperature (Tg) and solubility of the dendrimers were greatly inuenced by the terminal functional groups. Moore and coworkers3739 reported the preparation of hyperbranched aromatic poly(ether imide)s through the nucleophilic etherication of a protected AB2 monomer, as shown in Scheme 11. The self-polycondensation was carried out through the nucleophilic etherication of silylated phenol and aryl uoride in diphenylsulfone at 240 C in the presence of cesium uoride. A tert-butyldimethylsilyl (TBDMS) group was used to protect hydroxyl groups because TBDMS was less labile than trimethylsilyl groups and, therefore, allowed the thorough purication of the monomer by recrystallization. The number-average molecular weight (Mn), determined by GPC with light scattering calibration, ranged from 51,600 to 85,000. Transetherication occurred during the polymerization and resulted in an increase in Mw with the

Scheme 9. Melt polycondensation of AB2 monomers for the formation of hyperbranched aromatic polyamides.

1302

J. POLYM. SCI. PART A: POLYM. CHEM.: VOL. 42 (2004)

Scheme 10. Preparation of dendritic poly(ether imide) by stepwise reactions.

polymerization for 20 min.38 The transetherication also reduced DB of the hyperbranched poly(ether imide) from 66 (polymerization for 2.5 min) to 42% (polymerization for 20 min). The hyperbranched poly(ether imide)s were soluble in common organic solvents. The temperature for a 10% weight loss in air was 530 C, which implied that the polymer was one

of the most thermally stable hyperbranched polymers. Similarly to other hyperbranched polymers, the end groups of the hyperbranched polymers had a great inuence on the Tg value and solubility of the polymers. Moreover, the lm-forming ability and contact angle were also greatly inuenced by the end-group modication.39

Scheme 11. Preparation of hyperbranched poly(ether imide) by nucleophilic etherication.

HIGHLIGHT

1303

Figure 6. AB2-type monomers for hyperbranched poly(ether imide)s via nucleophilic etherication.

Other AB2-type monomers for hyperbranched poly(ether imide)s via nucleophilic etherication are listed in Figure 6.40 42 The effects of the end groups on the Tg value and solubility of the hyperbranched poly(ether imide)s are discussed in the literature. Commercially available linear poly(ether imide)s are produced from aromatic diamines and aromatic tetracarboxylic dianhydride via poly(amic acid) as a precursor polymer. We have reported the preparation of hyperbranched poly(ether imide) through the corresponding poly(amic acid) ester precursor, as shown in Scheme 12.43 45 The polymerization of the AB2 monomer prepared from 3,5-dimethoxyphenol and 4-nitrophthalonitrile proceeded in the presence of DBOP at room temperature. Subsequent chemical or thermal imidization of the resulting precursor gave a hyperbranched poly(ether imide) quantitatively. The hyperbranched polyimide terminated with acetyl amide had a DB of 0.48 and an Mw

value of 188,000, as determined by GPCMALLS. Differential scanning calorimetry (DSC) and thermogravimetric analysis measurements revealed a Tg of 193 C and a 10% weight-loss temperature of 470 C in air. The hyperbranched polyimide was soluble in aprotic polar solvents and had a low inherent viscosity. The hyperbranched polyimide lm with 4-methylphthalimide end groups possessed a lower dielectric constant, lower birefringence, and a shorter cutoff wavelength than the lm of poly(4,4-oxydiphenylene pyromellitimide), probably because of the looser packing structure of the hyperbranched polyimide. A hyperbranched poly(ether imide) without linear units was also prepared by an end-modication reaction of the precursor polymer with a branching agent.45 There are several publications concerning the preparation of hyperbranched polyimides via poly(amic acid) precursors from A2 and B3 monomers. Trifunctional

Scheme 12. Preparation of hyperbranched polyimides via poly(amic acid) precursors.

1304

J. POLYM. SCI. PART A: POLYM. CHEM.: VOL. 42 (2004)

Figure 7. Trifunctional monomers used as B3 monomers for hyperbranched poly(ether imide)s.

monomers used as B3 monomers for hyperbranched polyimides via poly(amic acid) precursors are listed in Figure 7. Fang et al.46,47 reported the preparation of hyperbranched polyimides from commercially available dianhydrides and a triamine monomer (20). The molar ratio and addition sequence of each component and the monomer concentration greatly affected the terminal structure of the resulting polymers and the gelation behavior of the polymerization. When a solution of 2,2bis(3,4-dicarboxyphenyl)hexauoropropane dianhydride (6FDA) was added to a solution of 20 in a molar ratio of 1:1, an amine-terminated hyperbranched polyimide was obtained (manner 1). However, the addition of a solution of 20 to a 6FDA solution in a molar ratio of 2:1 6FDA/20 yielded an anhydride-terminated hyperbranched polyimide (manner 2). In both cases, immediate gelation was observed when monomer solids were directly added or even if the addition of the monomer solution was not slow enough. It was found that the total solid content should be kept below 0.2 mol/L for manner 1 and below 0.075 mol/L for manner 2. A higher concentration led to an insoluble gel. Spectroscopic measurements supported the formation of the proposed hyperbranched polyimides and indicated DBs of 0.64 0.72. GPC measurements revealed that the hyperbranched polyimides had moderate Mns and broad molecular weight distributions. The gas permeation properties of membranes composed of the hyperbranched polyimides and a difunctional crosslinking agent were also examined.47 The amineterminated hyperbranched polyimides generally exhibited much higher gas permeability coefcients than the

corresponding anhydride-terminated ones. Triamine (21) was also used as a B3 monomer with commercially available dianhydrides (mainly 6FDA) to form hyperbranched poly(ether imide)s.48 50 End-group modication of the hyperbranched polyimides by photosensitive cinnamate or proton-conductive arenesulfonate groups was reported in the literature.49,50 We have reported the polycondensation of 1,4-phenylenediamine and tri(phthalic anhydride) (22) or tri(phthalic acid methyl ester) (23) in a 1:1 ratio to form hyperbranched polyimides via precursor polymers.51,52 Gelation could be avoided by the dropwise addition of a 1,4-phenylenediamine solution at 0 C to a solution of 22 at a total concentration of 0.017 g/mL (method A). A high reaction temperature, simultaneous addition, and a high polymerization concentration often resulted in gelation. High reactivity between anhydride and amino functional groups caused easy gel formation. However, the polymerization of 1,4-phenylenediamine and 23 in the presence of DBOP as a condensation agent was more controllable (method B). Gelation was avoided when the total concentration of the monomers was lower than 0.1 g/mL. This suggests that the in situ activation of the B3 molecule (method B) is suitable for A2 B3 polymerization for avoiding gelation. The inherent viscosity of the polymer prepared by method B was 0.17 0.97 and was dependent on the number of dendritic repeating units. Flexible and smooth polyimide lms that were transparent and yellow could be obtained from the precursor polymer prepared by method B. The high storage modulus of 3.1 4.0 GPa suggested the presence of good chain entanglement in the lms. Hyperbranched polyimides with the same repeating unit by the A2 B3 polymerization were prepared from an AB2 monomer (24; Fig. 8) to investigate the inuence of the starting monomers on the structure and properties of the resulting polyimides.52 The polymers prepared from the AB2 monomer had a compact highly branching structure, and

Figure 8. AB2 monomer for self-polycondensation for the formation of hyperbranched polyimides in the presence of DBOP.

HIGHLIGHT

1305

Scheme 13. Preparation of hyperbranched polyimides from A2 and BB2 monomers.

those prepared from the A2 and B3 monomers had a low branching density topology. Recently, the polymerization of 6FDA (A2) and 2,4,6triaminopyrimidine (TAP; BB2) for the formation of a hyperbranched polyimide was reported by Liu and Chung,53 as shown in Scheme 13. Because the 2-amino group in TAP had a lower reactivity than the 4- and 6-amino groups, no gelation was observed during the polymerization. The DB determined by 1H NMR measurements increased from 36 to 83% as the molar ratio of 6FDA to TAP increased from 1:1 to 2:1. It was reported that the hyperbranched polyimides showed good thermal stability, good solubility, and low inherent viscosity.

COPOLYMERIZATION OF AB2 AND AB MONOMERS


Generally, dendritic polymers show low solution viscosity because of the lack of chain entanglement, which might cause poor mechanical properties in the resulting polymers. In fact, cast lms prepared from hyperbranched polymers are usually brittle and hard to peel off as free-standing lms from a substrate. The copolymerization of AB2 and AB monomers results in the formation of hyperbranched copolymers with a controlled number of linear segments derived from the AB monomer. The effect of the branching or linear segments on the properties of hyperbranched copolymers can be quantitatively estimated by a copolymerization study.

We have reported the direct copolymerization of an AB2 monomer (13) and an AB monomer (25) for the formation of hyperbranched aromatic polyamide copolymers, as shown in Scheme 14.54 56 Compound 25 was prepared as an AB monomer; the reactivity of each functional group was presumed to be equal to that of 13. It was conrmed by 1H NMR measurements that the ratio of each monomer in the resulting copolymers was almost identical to the feed molar ratio of 13 and 25. Moreover, the ratio of ve kinds of possible repeating units for the copolymer (Fig. 9) was consistent with a statistical distribution.57 The inherent viscosity of the resulting copolymers increased as the feed ratio of the AB monomer increased. The feed molar ratio affected Tgs and softening points (Tss) of the resulting copolymers, as shown in Figure 10. A minimum at 197 C was observed for a composition with 50% AB2 monomer. Minimum peaks in the DSC measurements were also observed for aromatic polyester copolymers composed of linear and branching units.58,59 Tg for each composition was clearly lower than the value calculated by Foxs equation.60 There are two possible reasons for the Tg behavior of the copolymers. First, intermolecular entanglement in the copolymers might be reduced as the amount of the AB2 monomer increases, and this causes an increase in the free volume of the copolymer. Second, the number of unreacted B functions (amino groups) increases as the molar ratio of the AB2 monomer increases. It is not clear which one is crucial for the Tg behavior. The solubility of the hyperbranched

Scheme 14. Direct polymerization of AB2 and AB monomers for the formation of hyperbranched polyamide copolymers.

1306

J. POLYM. SCI. PART A: POLYM. CHEM.: VOL. 42 (2004)

Figure 9. Possible repeating units of the hyperbranched aromatic polyamide copolymer from 13 and 25.

polyamide copolymers is dependent on the feed molar ratio of 13 and 25. Although the linear polyamide prepared from 25 was only partially soluble in NMP, the resulting copolymers showed good solubility in organic solvents. It is surprising that the copolymerization of 13 (5 mol %) clearly improved the solubility of the resulting copolymer in comparison with the linear polymer prepared from 25. The improved solubility enabled lm preparation through the casting of the polymer solution onto a glass plate. The effect of the feed ratio on Youngs modulus, calculated by the initial slope of the stress strain curves, is shown in Figure 11.54 The modulus decreased with an increasing feed ratio of the AB2 monomer (13) and became almost constant over 60 mol % 13. The data suggest that the intermolecular entanglement of the copolymers was reduced by the incorporation of the AB2 branching unit. Other AB-type monomers examined as comonomers for the copolymerization with 13 or 16 are listed in Figure 12.56 The end-capping reaction of amino groups with benzoyl chloride reduced solute

Figure 11. Youngs modulus, determined by tensile testing, of the copolymer lms (13 and 25).

solute interactions and, therefore, improved the solubility of the copolymers. Moore and coworkers61 65 reported the preparation and properties of hyperbranched aromatic polyimide copolymers via nucleophilic substitution, as shown in Scheme 15. Copolymers with 0 1 molar fraction of the AB monomer (xAB) were prepared. Mw of the copolymers ranged from 59,300 to 115,000. DB of the copolymers was determined indirectly by (A B)/(A B2) competitive model compound reactions to be 0 0.67.63 The viscosity of the concentrated solution (10 wt % solids in NMP) of the copolymers increased rapidly when xAB was greater than 0.8. The intrinsic viscosities and MarkHouwink coefcients also increased noticeably in the same range. The calculated distance between branches (lAB) had a great inuence on the rheological properties. Continuous, free-standing lms that could be peeled off from a glass substrate were prepared for the copolymers with xAB 0.90. Tg of the copolymers increased gradually with xAB up to xAB 0.75 and then rose sharply toward 214 C for the linear polyimide. Moreover, the contact angle and surface energy of the copolymer lms were highly dependent on the sum of

Figure 10. (E) Tg and () Ts of the hyperbranched polyamide copolymers prepared from 13 and 25. The dashed line represents Tgs calculated with Foxs equation.60

Figure 12. AB-type monomers examined for hyperbranched polyamide copolymers.

HIGHLIGHT

1307

Scheme 15. Copolymerization of AB2 and AB monomers for the formation of hyperbranched polyimide copolymers.

the terminal segments derived from AB2 and AB monomers.64

CONCLUSIONS
Similarly to other dendritic macromolecules, the introduction of successive branching structures changes the properties of aromatic polyamides and polyimides dramatically. Improved synthetic pathways, such as an orthogonal/double-stage convergent approach and a convergent/divergent approach with unprotected AB2 building blocks, have been reported for dendritic aromatic polyamides. The improved pathways can reduce the number of reaction steps required for the preparation of high-generation dendrimers. Hyperbranched aromatic polyamides and polyimides [poly(ether imide)s] have been prepared from various ABx-type monomers. All of the hyperbranched polymers have shown good solubility and low solution viscosity despite rigid aromatic amide or imide units. End-group modication of the hyperbranched polymers greatly affects the properties, such as the solubility, Tg, lmforming ability, and contact angle. In addition to the self-polymerization of ABx-type monomers, the polymerization of A2 and B3 monomers seems to be an attractive route for producing hyperbranched polyamides and polyimides from practical viewpoint. The potential advantage of the A2 B3 polymerization lies on the use of commercially available monomers for the preparation of hyperbranched polymers. Improved polymerization systems with BB2 monomers for the A2 B3 polymerization approach have recently been reported for hyperbranched poly(amide imide)s and hyperbranched polyimides. The copolymerization of AB2 and AB monomers allows a quantitative investigation of the effect of the branching repeating units on the properties of the resulting copolymers. The tensile modulus of hyperbranched polyamide copolymers decreases as the feed ratio of the AB2 monomer increases and becomes almost constant over 60 mol %. The

rheological properties of hyperbranched polyimide copolymers are highly dependent on the calculated distance between branches. Both copolymerization studies suggest that a small number of AB2 branching units dramatically affects the properties of the resulting copolymers. In comparison with the number of publications on synthesis, there are only a few publications on the application of dendritic polyamides or polyimides. Hyperbranched aromatic polyamides have been examined as supports for protein immobilization and palladium nanoclusters.66 68 We have reported preliminary results for the preparation of composite materials of the hyperbranched aromatic polyamide derived from 13 with vinyl polymers and silica via a sol gel reaction.69,70 The application of hyperbranched poly(ether imide)s as photosensitive polymer matrices was reported by Ueda and coworkers.71,72 Blends of hyperbranched poly(ether imide) with polystyrene,73 polyolens,74 and bisphenol A based bis(maleimide)41 have also been reported in the literature. In addition to the stable and strong scaffold of dendritic polyamides and polyimides, the multivalence of dendritic polymers allows us to tune the properties of the nal products, such as the Tg, solubility, miscibility, and surface energy. Further extensive studies for the applications of dendritic polyamides and polyimides are required to broaden the eld of dendritic high-performance polymers.

REFERENCES AND NOTES


1. Newkome, G. R.; Mooreeld, C. N.; Vo gtle, F. Dendritic Molecules: Concepts, Syntheses, Perspectives; VCH: New York, 1996. 2. Dendrimers and Other Dendritic Polymers; Fre chet, J. M. J.; Tomalia, D. A., Eds.; Wiley: New York, 2001. 3. Frey, H. Angew Chem Int Ed Eng 1998, 37, 2193.

1308

J. POLYM. SCI. PART A: POLYM. CHEM.: VOL. 42 (2004) 33. Yang, G.; Jikei, M.; Kakimoto, M. Macromolecules 1998, 31, 5964. 34. Reichert, V. R.; Mathias, L. J. Macromolecules 1994, 27, 7024. 35. Leu, C.-M.; Chang, Y.-T.; Shu, C.-F.; Teng, C.-F.; Shiea, J. Macromolecules 2000, 33, 2855. 36. Leu, C.-M.; Shu, C.-F.; Teng, C.-F.; Shiea, J. Polymer 2001, 42, 2339. 37. Thompson, D. S.; Markoski, L. J.; Moore, J. S. Macromolecules 1999, 32, 4764. 38. Thompson, D. S.; Markoski, L. J.; Moore, J. S.; Sendijarevic, I.; Lee, A.; McHugh, A. J. Macromolecules 2000, 33, 6412. 39. Orlicki, J. A.; Thompson, D. S.; Markoski, L. J.; Sill, K. N.; Moore, J. S. J Polym Sci Part A: Polym Chem 2002, 40, 926. 40. Wu, F.-I.; Shu, C.-F. J Polym Sci Part A: Polym Chem 2001, 39, 2536. 41. Baek, J.-B.; Qin, H.; Mather, P. T.; Tan, L.-S. Macromolecules 2002, 35, 4951. 42. Li, X.; Li, Y.; Tong, Y.; Shi, L.; Liu, X. Macromolecules 2003, 36, 5537. 43. Yamanaka, K.; Jikei, M.; Kakimoto, M. Macromolecules 2000, 33, 1111. 44. Yamanaka, K.; Jikei, M.; Kakimoto, M. Macromolecules 2000, 33, 6937. 45. Yamanaka, K.; Jikei, M.; Kakimoto, M. Macromolecules 2001, 34, 3910. 46. Fang, J.; Kita, H.; Okamoto, K. Macromolecules 2000, 33, 4639. 47. Fang, J.; Kita, H.; Okamoto, K. J Membr Sci 2001, 182, 245. 48. Chen, H.; Yin, J. J Polym Sci Part A: Polym Chem 2002, 40, 3804. 49. Chen, H.; Yin, J. Polym Bull 2003, 49, 313. 50. Chen, H.; Yin, J.; Xu, H. Polym J 2003, 35, 280. 51. Hao, J.; Jikei, M.; Kakimoto, M. Macromolecules 2002, 35, 5372. 52. Hao, J.; Jikei, M.; Kakimoto, M. Macromolecules 2003, 36, 3519. 53. Liu, Y.; Chung, T.-S. J Polym Sci Part A: Polym Chem 2002, 40, 4563. 54. Jikei, M.; Fujii, K.; Yang, G.; Kakimoto, M. Macromolecules 2000, 33, 6228. 55. Jikei, M.; Fujii, K.; Kakimoto, M. J Polym Sci Part A: Polym Chem 2001, 39, 3304. 56. Jikei, M.; Fujii, K.; Kakimoto, M. Macromol Symp 2003, 199, 223. 57. Frey, H.; Ho lter, D. Acta Polym 1999, 50, 67. 58. Kricheldorf, H. R.; Sto ber, O. S.; Lu bbers, D. Macromol Chem Phys 1995, 196, 3549. 59. Kricheldorf, H. R.; Stukenbrock, T. J Polym Sci Part A: Polym Chem 1998, 36, 2347. 60. 1/Tg Wa/Tga Wb/Tgb. See Fox, D. W.; Allen, R. B. In Encyclopedia of Polymer Science and Engineering; Mark, H. F.; Bikales, N. M.; Overberger, C. G.; Menges, G., Eds.; Wiley: New York, 1988; Vol. 3, p 758.

4. Malmstro m, E.; Hult, A. J Macromol Sci Rev Macromol Chem Phys 1997, C37, 555. 5. Kim, Y. H. J Polym Sci Part A: Polym Chem 1998, 36, 1685. 6. Voit, B. J Polym Sci Part A: Polym Chem 2000, 38, 2505. 7. Jikei, M.; Kakimoto, M. Prog Polym Sci 2001, 26, 1233. 8. Miller, T. M.; Neenan, T. X. Chem Mater 1990, 2, 346. 9. Bayliff, P. M.; Feast, W. J.; Parker, D. Polym Bull 1992, 29, 265. 10. Backson, S. C. E.; Bayliff, P. M.; Feast, W. J.; Kenwright, A. M.; Parker, D.; Richards, R. W. Macromol Symp 1994, 77, 1. 11. Ishida, Y.; Jikei, M.; Kakimoto, M. Macromolecules 2000, 33, 3202. 12. Yamakawa, Y.; Ueda, M.; Takeuchi, K.; Asai, M. Macromolecules 1999, 32, 8363. 13. Hayakawa, T.; Yamakawa, Y.; Nomura, M.; Okazaki, M.; Takeuchi, K.; Asai, M.; Ueda, M. Polym J 2000, 32, 784. 14. Okazaki, M.; Washio, I.; Shibasaki, Y.; Ueda, M. J Am Chem Soc 2003, 125, 8120. 15. Kim, Y. H. J Am Chem Soc 1992, 114, 4947. 16. Yang, G.; Jikei, M.; Kakimoto, M. Proc Jpn Acad B 1998, 74, 188. 17. Yang, G.; Jikei, M.; Kakimoto, M. Macromolecules 1999, 32, 2215. 18. Russo, S.; Boulares, A.; Mariani, A. Macromol Symp 1998, 128, 13. 19. Ishida, Y.; Sun, A. C. F.; Jikei, M.; Kakimoto, M. Macromolecules 2000, 33, 2832. 20. Haba, O.; Tajima, H.; Ueda, M.; Nagahata, R. Chem Lett 1998, 333. 21. Monticelli, O.; Mendichi, R.; Bisbano, S.; Mariani, A.; Russo, S. Macromol Chem Phys 2000, 201, 2123. 22. Hawker, C. J.; Lee, R.; Fre chet, J. M. J. J Am Chem Soc 1991, 113, 4583. 23. Ho lter, D.; Burgath, A.; Frey, H. Acta Polym 1997, 48, 30. 24. Flory, P. J. Principles of Polymer Chemistry; Cornell University Press: Ithaca, NY, 1953; Chapter 9. 25. Aharoni, S. M.; Edwards, S. F. Macromolecules 1989, 22, 3361. 26. Aharoni, S. M.; Murthy, N. S.; Zero, K.; Edwards, S. F. Macromolecules 1990, 23, 2533. 27. Aharoni, S. M. Macromolecules 1991, 24, 235. 28. Aharoni, S. M. Polym Adv Technol 1995, 6, 373. 29. Jikei, M.; Chon, S.-H.; Kakimoto, M.; Kawauchi, S.; Imase, T.; Watanabe, J. Macromolecules 1999, 32, 2061. 30. Russo, S.; Boulares, A.; Mariani, A.; Cosulich, M. E. Macromol Symp 1999, 143, 309. 31. Komber, H.; Voit, B.; Monticelli, O.; Russo, S. Macromolecules 2001, 34, 5487. 32. Chang, Y.-T.; Shu, C.-F. Macromolecules 2003, 36, 661.

HIGHLIGHT 61. Markoski, L. J.; Thompson, D. S.; Moore, J. S. Macromolecules 2000, 33, 5315. 62. Markoski, L. J.; Moore, J. S.; Sendijarevic, I.; McHugh, A. J. Macromolecules 2001, 34, 2695. 63. Markoski, L. J.; Thompson, D. S.; Moore, J. S. Macromolecules 2002, 35, 1599. 64. Orlicki, J. A.; Viernes, N. O. L.; Moore, J. S.; Sendijarevic, I.; McHugh, A. J. Langmuir 2002, 18, 9990. 65. Sendijarevic, I.; Liberatore, M. W.; McHugh, A. J.; Markoski, L. J.; Moore, J. S. J Rheol 2001, 45, 1245. 66. Cosulich, M. E.; Russo, S.; Pasquale, S.; Mariani, A. Polymer 2000, 41, 4951. 67. Tabuani, D.; Monticelli, O.; Komber, H.; Russo, S. Macromol Chem Phys 2003, 204, 1576.

1309

68. Tabuani, D.; Monticelli, O.; Chincarini, A.; Bianchini, C.; Vizza, F.; Moneti, S.; Russo, S. Macromolecules 2003, 36, 4294. 69. Kaneko, R.; Jikei, M.; Kakimoto, M. High Perform Polym 2002, 14, 41. 70. Kaneko, R.; Jikei, M.; Kakimoto, M. High Perform Polym 2002, 14, 105. 71. Okazaki, M.; Shibasaki, Y.; Ueda, M. Chem Lett 2001, 762. 72. Okazaki, M.; Shibasaki, Y.; Ueda, M. J Photopolym Sci Tech 2001, 14, 45. 73. Orlicki, J. A.; Moore, J. S.; Sendijarevic, I.; McHugh, A. J. Langmuir 2002, 18, 9985. 74. Sendijarevic, I.; McHugh, A. J.; Orlicki, J. A.; Moore, J. S. Polym Eng Sci 2002, 42, 2393.

Вам также может понравиться