Sunteți pe pagina 1din 8

Nucleobase-Induced Supramolecular Polymerization in the Solid State

STUART J. ROWAN, PHIRIYATORN SUWANMALA, SONA SIVAKOVA W. M. Keck Laboratories for Organic Synthesis, Department of Macromolecular Science and Engineering, Case Western Reserve University, 2100 Adelbert Road, Cleveland, Ohio 44106

Received 7 June 2003; accepted 27 August 2003

ABSTRACT: Supramolecular polymerization, that is, the self-assembly of polymer-like materials through the utilization of the noncovalent bond, has been a developing area of research over the last decade. In this article, we report the synthesis of nucleobaseterminated (N6-anisoyl-adenine and thymine) low-molecular-weight poly(tetrahydrofuran) macromonomers (2000 g mol1). The adenine-derived supramolecular telechelic polymer self-assembled in the solid state to yield materials with lm- and ber-forming capabilities. This material was thermally reversible and exhibited a ceiling temperature, above which a drop in viscosity was observed and bers could no longer be obtained. 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 3589 3596, 2003 Keywords: nucleobase; macromonomers; reversible polymers; supramolecular structures; self-assembly

INTRODUCTION
Supramolecular polymerization,13 that is, the self-assembly of polymer-like materials through the utilization of the noncovalent bond, has been a developing area of research over the last decade, with a number of groups investigating how different facets of noncovalent chemistry can be used to induce the formation of polymeric aggregates4 13 and nanoparticles.14,15 Such supramolecular polymers potentially have a number of interesting properties: (1) they form spontaneously, without the need for an initiation process (or catalyst); (2) they are dynamic, that is, they are formed under reversible conditions, thus imparting the potential of such materials to be thermally selfrepairing;16,17 and (3) termination processes during the polymerization (self-assembly) are limited. As a result, the degree of polymerization depends, to a large extent, on the strength of the
Correspondence to: S. J. Rowan (E-mail: sjr4@po.cwru.edu)
Journal of Polymer Science: Part A: Polymer Chemistry, Vol. 41, 3589 3596 (2003) 2003 Wiley Periodicals, Inc.

supramolecular interaction between the monomers (Ka) and the monomer concentration [M]. If growth of the supramolecular polymer operates through a multistage open-association mechanism, where the binding constant is independent of molecular weight, then the degree of polymerization will be proportional to (Ka[M])1/2. As such, large values of Ka are generally required to obtain aggregates of signicant molecular weight. For monomer units that exhibit degrees of interaction generally too low to form polymers through the multistage open-association mechanism, other mechanisms have been used to induce these molecules to form polymer aggregates. For example, one way to increase the degree of polymerization is to use the anisotropy of a liquid-crystalline environment to enhance the degree of interaction between the monomers.18 21 Our initial investigations in this eld have focused on the use of nucleobase pair interactions to control the selfassembly of macromonomers into polymeric architectures.2225 We report the synthesis of complementary supramolecular telechelic polymers that have one nucleobase either thymine (T) or
3589

3590

ROWAN, SUWANMALA, AND SIVAKOVA

N6-(4-methoxybenzoyl)adenine (AAn)at each chain end and the effect that such a synthetic modication has on the properties of the material.

EXPERIMENTAL
Materials Thymine acetic acid and N6-(4-methoxybenzoyl)9-(carboxymethyl)adenine were synthesized according to literature procedures.26,27 All reagents and solvents were purchased from Aldrich Chemical Co. Reagents were used without further purication. Solvents were distilled from suitable drying agents. Characterization NMR spectra were recorded on either a Varian Gemini 200-MHz NMR or a Varian 600-MHz spectrometer. Differential scanning calorimetry (DSC) experiments were carried out on a Pyris 1 DSC (PerkinElmer) under owing N2 at a temperate rate of 5 C/min. Fiber and lm images were obtained with an optical microscope (Olympus BX 60). Fourier transform infrared (FTIR) measurements were performed on a ABB Bomen MB Series FTIR spectrometer. The molecular weights of the polymers were measured by mass spectrometry on a Bruker BIFLEX III matrixassisted laser desorption/ionization time-of-ight mass spectrometer (MALDI-TOF MS). Preparation of Thymine-Terminated Poly(tetrahydrofuran) (T2T) To 1.00 g (5.43 mmol) of thymine acetic acid (3T) was added 10 mL of dry dimethylformamide (DMF). The solution was stirred for 30 min while cooling in an ice bath. Trimethylacetyl chloride (0.50 mL, 4.06 mmol) was added dropwise, and the reaction was stirred for an additional 10 min. N-Methyl morpholine (1.12 mL, 10.2 mmol) was added dropwise with a syringe, and the reaction was stirred for another hour. To the stirring solution was added 1.50 g (1.36 mmol) of bis(3aminopropyl)-terminated poly(tetrahydrofuran) (1) dissolved in 20 mL of dry DMF. The reaction was allowed to warm to room temperature and stirred for 48 h. The solvent was removed, and the residue was treated with ethyl acetate. The resulting precipitate was ltered off and dried in high vacuum. This crude product was reprecipitated from a mixture of water and methanol

(MeOH) to give a white solid product in 1.45 g (74%) yield. This reaction procedure was repeated three times to ensure complete reaction of the amino chain ends. 1 H NMR [600 MHz, deuterated dimethyl sulfoxide (DMSO-d6)]: 11.20 (2H, s, thymine NH), 8.05 (2H, br s, NH), 7.37 (2H, s, thymine H-6), 4.20 (4H, s, thymine CH2), 3.32 (48H, m, CH2), 3.08 (4H, m, CH2), 1.71 (6H, s, CH3), 1.58 (4H, m, CH2), 1.46 (44H, m, CH2). 13C NMR (50 MHz, DMSO-d6): 12.3, 26.5, 29.6, 36.3, 49.7, 67.8, 70.1, 108.3, 142.7, 151.4, 164.8, 167.0. FTIR [annealed (140 C) lm]: 3316, 2939, 2854, 1722, 1665, 1559, 1481, 1429, 1337, 1236, 1173, 1110 cm1. MALDI-TOF MS [2-(hydroxyphenylazo)benzoic acid)]: number-average molecular weight (Mn) 1577, weight-average molecular weight (Mw) 1750 g/mol, polydispersity (PDI) 1.11; DSC [melting temperature (Tm) 155 C]. Preparation of N6-Protected Adenine-Terminated Poly(tetrahydrofuran) (AAn3AAn) N6-(4-Methoxybenzoyl)-9-(carboxymethyl)adenine (3AAn) (0.60 g, 1.83 mmol) was dissolved in dry DMF (10 mL). The solution was cooled in an ice bath and stirred for 30 min. Trimethylacetyl chloride (0.17 mL, 1.38 mmol) was added dropwise with a syringe, and the reaction was stirred for an additional 10 min. N-Methyl morpholine (0.38 mL, 3.46 mmol) was added dropwise, and then the reaction was stirred for another hour. To the stirring solution was added 0.50 g (0.46 mmol) of bis(3-aminopropyl)-terminated 1 dissolved in 20 mL of dry DMF. The reaction was stirred for 48 h while allowing to warm to room temperature. The solvent was removed, and the residue was treated with EtOAc. The resulting precipitate was ltered off and washed thoroughly with MeOH to give a light yellow solid product. The material was further puried by passing through a silica plug with CH2Cl2/MeOH (100:0, 99:1,. . ., 90:10) to yield an off-white powder (0.40 g, yield: 51%). 1 H NMR (600 MHz, DMSO-d6): 10.97 (2H, br s, adenine NH), 8.65 (2H, s, adenine H-8), 8.35 (4H, s, adenine H-2, and NH), 8.02 (4H, d, ArH), 7.04 (4H, d, ArH), 4.92 (4H, s, adenine CH2), 3.82 (6H, s, ArOCH3), 3.31 (40H, m, CH2), 3.12 (4H, s, CH2), 1.63 (4H, m, CH2), 1.46 (36H, m, CH2). 13C NMR (50 MHz, DMSO-d6): 26.4, 29.6, 36.5, 45.6, 55.8, 67.8, 7.1, 114.0, 125.2, 125.9, 130.9, 145.7, 150.6, 151.7, 152.9, 162.9, 166.3. FTIR [annealed (140 C) lm]: 3298, 3087, 2942, 2854, 1677, 1611, 1580, 1487, 1457, 1315, 1251, 1176, 1109, 1034

SUPRAMOLECULAR POLYMERIZATION

3591

Scheme 1. Synthesis of the supramolecular telechelic macromonomers AAn2AAn and T2T.

cm1. MALDI-TOF MS (-cyano-4-hydroxycinnamic acid): Mn 1719, Mw 1953 g/mol, PDI 1.14; DSC (Tm 134 C). Binding Studies The binding studies were performed with deuterated chloroform predried with potassium carbonate. To a stock solution of dodecyl-AAn (5 mM) was titrated a mixture of dodecyl-T (20 mM) and dodecyl-AAn (5 mM) so that the concentration of dodecyl-T varied between 1 and 16 mM. The shift of the hydrogen on the N6 position of the adenine was monitored with respect to chloroform, and the data were tted to the equation for a 1:1 binding.

RESULTS AND DISCUSSION


Design, Synthesis, and Characterization of the Nucleobase-Terminated Poly(tetrahydrofuran) Purines, such as adenine and guanine, have two possible binding sites, one on the WatsonCrick face and one on the Hoogsteen face. In an attempt to limit the binding possibilities of the adenine, we decided to use the N6-anisoyl-protected adenine as the binding motif for these studies. Although such a synthetic modication may reduce the degree of binding of adenine,28 an amide group will tend to block the WatsonCrick face of the nucleobase. We measured the binding constant in CDCl3 between the model compounds dodecyl-T and dodecyl-AAn to be about 21 M1, by NMR titration experiments and monitoring the shift of the N6-H. Although this conrmed that the AAn and T still interact with each other, it

does correspond to a reduction of about ve times in the binding constant as compared with the literature value of the A-T interaction (CDCl3 KA-T 100 M1).29 The low-molecular-weight macromonomer to which we chose to attach the nucleobase derivatives was the commercially available bis(3-aminopropyl)-terminated poly(tetrahydrofuran) (1, Mn ca. 1400). The synthesis of the desired supramolecular telechelic macromonomers BP2BP (Scheme 1) was achieved by reacting the appropriate acetic acid nucleobase derivative 3BP [either N6-(4-methoxybenzoyl)-9-(carboxymethyl)adenine (3AAn) or thymine acetic acid (3T)] with 1 under mixed anhydride peptide coupling conditions. This reaction was repeated three times to ensure complete conversion of the amino chain ends. The structure of the resulting materials was conrmed by both NMR and MALDI-TOF MS. The 600-MHz 1H NMR of both T2T and AAn2AAn in DMSO-d6 are illustrated in Figure 1(a,b). The peaks for the end groups can be easily identied, and the corresponding molecular weights (Mns) of these materials can be calculated to be about 1300 and 1400 gmol1, respectively. The MALDI-TOF MS spectrum of AAn2AAn is depicted in Figure 2, and the molecular weight (Mn) corresponds to about 1700 gmol1 with a PDI of about 1.14. MALDI-TOF MS also conrmed the absence of any monosubstituted species in the sample. The small peaks observed between the major [M H] peaks correspond to [M Na] and [M K]. Table 1 lists the molecular weight and thermal characteristics of the two nucleobase-terminated poly(tetrahydrofuran)s along with the starting material 1 and the bis-acetylated material H2H, which was prepared as a control. The underivatized 1 was a soft waxy material with a melting

3592

ROWAN, SUWANMALA, AND SIVAKOVA

Figure 1. 1H NMR (600 MHz, DMSO-d6) of the nucleobase-terminated polymers (a) T2T and (b) AAn2AAn.

point of around 21 C. Placement of the nucleobase derivatives on the ends of 1 resulted in a marked change in the properties of the material. For example, an increase of over 100 C in the melting point was observed upon such derivatiza-

tion. A similar increase was not observed when the macromonomer was derivatized with acetate groups, and the resulting material was an oil at room temperature (melting temperature of about 20 C by DSC). This suggested the marked change

SUPRAMOLECULAR POLYMERIZATION

3593

Figure 2. MALDI-TOF MS of the nucleobase-terminated polymer AAn2AAn.

in properties of these materials is mainly because of the presence of the nucleobases and not only the presence of the amide groups used to link the nucleobase to the core. Although T2T exhibited both melting (155 C) and crystallization (145 C) points upon heating and subsequent cooling, the DSC thermograms for the AAn2AAn macromonomer exhibited three endotherms at 83, 106, and 134 C only on the rst heating [Fig. 3(a,b)]. Further cooling and heating cycles revealed only a small transition at about 22 C. Moreover, the AAn2AAn, which was

obtained from the reaction as a white powder, was converted (134 C) into a clear melt from which both bers [Fig. 4(a)] and lms [Fig. 4(b)] could be obtained. This enhancement of mechanical stability was consistent with the fact that this macromonomer self-assembles into higher molecular weight supramolecular structures. The lms of AAn2AAn appeared optically clear and showed no melting point in the DSC, although the transition at 22 C was observed. Solid-state FTIR of AAn2AAn before and after annealing provided some evidence for the formation of H bonding in

Table 1. Properties of the Bis-Functionalized Poly(tetrahydrofuran) Derivatives Mna (g mol1) 1400 1800 1600 1700 Tm (C) 21 20 155c 134d 113d 149c Fiber-Forming Temperature Rangeb No bers No bers No bers 104130 C 99123 C No bers

Macromonomer 1 H2H T2T AAn2AAn AAn2AAn:T2T (9:1) AAn2AAn:T2T (1:1)


a b

PDa 1.29 1.22 1.11 1.14

Calculated from MALDI-MS data. Temperature range at which bers can be obtained from the melt. c DSC second heating, 5 C/min. d DSC rst heating, 5 C/min; no major peaks were observed on reheating.

3594

ROWAN, SUWANMALA, AND SIVAKOVA

these peaks shifted to yield two peaks at lower wave numbers (3298 and 3087 cm1), indicative of the presence of hydrogen bonding. In addition, the two carbonyl peaks in the preannealed sample [1695 (adenine-amide) and 1659 cm1 (alkylamide)] converged into one broader peak centered at 1677 cm1. The shift of the amide on the adenine to lower wave numbers was consistent with its participation in H bonding. The ability to form lms and bers was not observed in the amineterminated 1, acetylated H2H, nor T2T, suggesting that it is the N6-4-methoxybenzoyladenine moiety that causes this behavior. Effect of Temperature on the Materials If supramolecular polymers are being formed in the solid state, then the material properties of these compounds would be expected to be sensitive to temperature. A rise in temperature should result in a weakening of the noncovalent bonds holding the polymer together, leading to a subsequent depolymerization of the material. This would effectively be the ceiling temperature (Tc) of the polymer. Although AAn2AAn melts at 134 C upon the rst heating, the material does not solidify upon cooling until about 104 C. Above this lower limit, bers could be pulled from the melt until the temperature reaches 130 C, at which point the viscosity of the material visually drops and bers can no longer be obtained. There are number of questions that arise upon observing this behavior. For example, why can lms and bers be formed by AAn2AAn and not by T2T? This question arises because the homodimerization of both nucleobases is weak (thymine and adenine homodimerization were calcu-

Figure 3. DSC heating and cooling curves of (a) T2T and (b) AAn2AAn.

this system. A number of broad peaks were observed in the NH region of the spectrum at 3470, 3358, 3301, and 3102 cm1. Upon annealing,

Figure 4. (a) Optical microscopy (100) of a ber of AAn2AAn obtained from the melt and (b) a photograph of a transparent lm of AAn2AAn.

SUPRAMOLECULAR POLYMERIZATION

3595

lated to be about 3.5 and 2.4 M1, respectively, in CDCl3).29 Furthermore, placement of the protecting group onto adenine should further lower this degree of interaction.28 From NMR dilution studies of N6-(4-methoxybenzoyl)-9-(dodecyl)adenine (dodecyl-AAn), we estimated the dimerization constant of N6-anisoyl adenine to be very low (3 M1 in CDCl3). Therefore, there is probably another self-assembling driving force taking place in this system that aids the supramolecular polymerization process in the solid state. We believe other factors such as - stacking, dipole dipole interactions, phase segregation30 34 between the hard nucleobase components, and the soft poly(tetrahydrofuran) chain35,36 and hydrogen bonding of the amide bond, which links the nucleobase to the polymer, may also aid the self-assembly process, and as such allow the formation of lms and bers from the melt. We believe that the presence of the high crystallization temperature (ca. 145 C) combined with a reduced surface (compared to AAn) may explain why T2T exhibits no ber-forming ability. We suspect that when the thymine derivative melts it is already above its Tc and does not form supramolecular polymers in the melt state. To further investigate these systems, we added the complementary non-berforming macromonomer T2T to AAn2AAn. Mixing of T2T and AAn2AAn in a 1:1 ratio resulted in a material from which the FTIR of the annealed lms showed broad peaks of about 3300 and 1665 cm1 consistent with H bonding being present in this system. This material did not melt until 149 C, and at this temperature no bers could be obtained. However, bers were obtainable when only 10% T2T was mixed with AAn2AAn, and this system exhibited a ber-forming temperature range of 99 123 C. Addition of a Supramolecular Chain-Terminating Agent We believe that the increase in mechanical properties in the AAn2AAn system was due to the presence of physical crosslinking, caused by phase segregation, and not simply the formation of long linear high-molecular-weight chains. To further test this theory, we added a small percentage of a supramolecular chain terminator. If only linear supramolecular chains are formed, then the addition of even a small amount of a monosubstituted adenine derivative should result in a depolymerization of the aggregate37 and consequently a reduction in the materials ability to

form bers. Dodecyl-AAn was, therefore, added to AAn2AAn, and the resulting material was examined. Although the addition of 50% of dodecyl-AAn does eliminate the ber-forming ability of AAn2AAn, the addition of smaller quantities of the chain terminator (up to 20%) did not. However, weaker, more brittle bers were obtained from such mixtures, and the Tc of the material dropped slightly from 130 to 116 C. The fact that bers could be obtained at all with the addition of such large quantities of dodecyl-AAn is consistent with the presence of some form of noncovalent or physical crosslinking, for example, phase segregation, being present in the melt of this system, which helps to increase the effective molecular weight of the sample, even in the presence of a chain terminator.

CONCLUSIONS
We attached thymine and adenine derivatives to low-molecular-weight poly(tetrahydrofuran) and found that the addition of the N6-(4-methoxybenzoyl)adenine moiety to the chain ends of the macromonomer conferred polymer-like properties, such as lm and ber formation, on this otherwise low-molecular-weight (2000 g mol1) material. Investigations are now under way to further understand the mechanical and solution properties of these supramolecular materials. In addition, we are also looking at the role the N6-adenine protecting group has, if any, on this self-assembly process.
The authors thank the Royal Thai Government, the Dow Chemical Co., and Case Western Reserve University for funding and J. Ben Beck for obtaining the NMR data.

REFERENCES AND NOTES


1. Brunsveld, L.; Folmer, B. J. B.; Meijer, E. W.; Sijbesma, R. P. Chem Rev 2001, 101, 4071 4097. 2. Ciferri, A. Macromol Rapid Commun 2002, 23, 511529. 3. Supramolecular Polymers; Credi, A., Ed.; Marcel Dekker: New York, 2000. 4. Zimmerman, S. C.; Zeng, F. W.; Reichert, D. E. C.; Kolotuchin, S. V. Science 1996, 271, 10951098. 5. Castellano, R. K.; Nuckolls, C.; Eichhorn, S. H.; Wood, M. R.; Lovinger, A. J.; Rebek, J., Jr. Angew Chem Int Ed Engl 1999, 38, 26032606.

3596

ROWAN, SUWANMALA, AND SIVAKOVA

6. Yamaguchi, N.; Gibson, H. W. Angew Chem Int Ed Engl 1999, 38, 143147. 7. Zhang, P.; Moore, J. S. J Polym Sci Part A: Polym Chem 2000, 38, 207219. 8. Zubarev, E. R.; Pralle, M. U.; Sone, E. D.; Stupp, S. I. J Am Chem Soc 2001, 123, 4105 4106. 9. Archer, E. A.; Krische, M. J. J Am Chem Soc 2002, 124, 5074 5083. 10. Sijbesma, R. P.; Beijer, F. H.; Brunsveld, L.; Folmer, B. J. B.; Hirschberg, J. H. K. K.; Lange, R. F. M.; Lowe, J. K. L.; Meijer, E. W. Science 1997, 278, 16011604. 11. Hirschberg, J. H. K. K.; Beijer, F. H.; van Aert, H. A.; Magusin, P. C. M. M.; Sijbesma, R. P.; Meijer, E. W. Macromolecules 1999, 32, 2696 2705. 12. Lohmeijer, B. G. G.; Schubert, U. S. J Polym Sci Part A: Polym Chem 2003, 41, 11431427. 13. El-Ghayoury, A.; Schenning, A. P. H. J.; Meijer, E. W. J Polym Sci Part A: Polym Chem 2002, 40, 4020 4023. 14. Becker, M. L.; Remsen, E. E.; Wooley, K. L. J Polym Sci Part A: Polym Chem 2001, 39, 4152 4166. 15. Manners, I. J Polym Sci Part A: Polym Chem 2002, 40, 179 191. 16. Chen, X.; Dam, M. A.; Ono, K.; Mal, A.; Shen, H.; Nut, S. R.; Sheran, K.; Wudl, F. Science 2002, 295, 1698 1702. 17. For an example of a self-healing material that operates with a different mechanism, see: White, S. R.; Sottos, N. R.; Geubelle, P. H.; Moore, J. S.; Kessler, M. R.; Sriram, S. R.; Brown, E. N.; Viswanathan, S. Nature 2001, 409, 794 797. 18. Bladon, P.; Grifn, A. C. Macromolecules 1993, 26, 6604 6610. 19. Kotera, M.; Lehn, J.-M.; Vigneron, J.-P. J Chem Soc Chem Commun 1994, 197199.

20. Kotera, M.; Lehn, J.-M.; Vigneron, J.-P. Tetrahedron 1995, 51, 19531972. 21. St. Pourcain, C.; Grifn, A. C. Macromolecules 1995, 28, 4116 4121. 22. Shimizu, T.; Iwaura, R.; Masuda, M.; Hanada, T.; Yase, K. J Am Chem Soc 2001, 123, 59475955. 23. Yamaguchi, K.; Lizotte, J. R.; Long, T. E. Macromolecules 2002, 35, 8745 8750. 24. Iwaura, R.; Yoshida, K.; Masuda, M.; Yase, K.; Shimizu, T. Chem Mater 2002, 14, 30473053. 25. Fogleman, E. A.; Yount, W. C.; Xu, J.; Craig, S. L. Angew Chem Int Ed Engl 2002, 41, 4026 4028. 26. Kosynkina, L.; Wei Wang, T.; Liang, C. Tetrahedron Lett 1994, 35, 51735176. 27. Uhlmann, E.; Will, D. W.; Breipohl, G.; Langner, D.; Knolle, J. Tetrahedron 1995, 51, 12069 12082. 28. Nowick, J. S.; Chen, J. S.; Noronha, G. J Am Chem Soc 1993, 115, 7636 7644. 29. Sartorius, J.; Schneider, H.-J. Chem Eur J 1996, 2, 1446 1452. 30. Fo rster, S.; Plantenberg, T. Angew Chem Int Ed Engl 2002, 41, 688 714. 31. Stupp, S. I. Curr Opin Colloid Interface Sci 1998, 3, 20 26. 32. Balsamo, V.; Gyldenfeldt, F.; Stadler, R. Macromol Chem Phys 1996, 197, 33173341. 33. de Fucca Freitas, L.; Jacobi, M. M.; Gonc alves, G. Macromolecules 1998, 31, 3379 3382. 34. Abed, S.; Boileau, S.; Bouteiller, L. Macromolecules 2000, 33, 8479 8487. 35. Lillya, C. P.; Baker, R. J.; Hu tte, S.; Winter, H. H.; Lin, Y.-G.; Shi, J.; Dickinson, C.; Chien, J. C. W. Macromolecules 1992, 25, 2076 2080. jelund, K.; Loontjens, T.; Steeman, P.; Palmans, A.; 36. O Maurer, F. Macromol Chem Phys 2003, 204, 52 60. 37. Folmer, B. J. B.; Cavini, E.; Sijbesma, R. P.; Meijer, E. W. Chem Commun 1998, 18471848.

S-ar putea să vă placă și