Вы находитесь на странице: 1из 7

Self-Assembly

of Organic Nano-Objects into Functional Materials

Samuel I. Stupp, Martin U. Pralle, Gregory N. Tew, Leiming Li, Mehmet Sayar, and Eugene R. Zubarev
Introduction
One of the goals of contemporary science is the atomic or molecular design of materials in order to achieve specific properties. There is special interest in imitating with these designed materials the remarkable integrated functionality we see in biology. Soft matter offers a particularly good opportunity to realize these goals because of the vast structural space offered by organic systems. One subset of designed organic materials is that in which the constituent units are groups of molecules that achieve a specific shape and size (see Figure 1). We refer to these systems as supramolecular materials, since their rational synthesis goes beyond molecular structure and requires supramolecular chemistry. This branch of chemistry aims to control intermolecular, noncovalent interactions in the same way that synthetic organic chemists control the formation of covalent bonds.13 One could argue that common semicrystalline polymers, liquid crystals, and monolayers are supramolecular because groups of molecules in such systems are, after all, defined by defects or external boundaries. The materials of interest in this article are composed of molecular aggregates and are therefore inherently supramolecular in nature. In these systems, thermodynamically or kinetically defined defects and surfaces are not the controlling factors of supramolecular structure. Aggregate formation is controlled instead by molecular architecture. Supramolecular materials composed of large and regular molecular clusters can dimensionally amplify the lattices we know in crystalline materials into superlattices with large periodicities. This also suggests that efficient packing considerations would affect symmetry preferences and allow one to predict structure at

longer length scales than those characteristic of the clusters themselves. Good examples would include the tendency of flat, platelike groups of molecules to stack parallel to each other and the tendency of tubes or cylindrical structures to align along a common direction. This latter feature of objects with high aspect ratios is well known in liquid-crystalline phases.4 Our laboratory showed a few years ago that chiral and chemically reactive molecules could be designed to self-assemble into two-dimensional structures. In these structures, defects control the x-y dimensions of the system and molecular design determines the z coordinate (see Figure 2).5,6 Furthermore, the reactivity of these molecules made the layered assemblies convertible to two-dimensional polymers, a set of molecular objects predicted theoretically to have many interesting properties.712 The properties we found in such systems include the thermal stability of stacked plates5 and the spontaneous formation of nanoporous structures as a result of poor tiling among irregular twodimensional objects.13 In supramolecular form, these materials are examples in which defects control the boundaries of molecular aggregates. We describe in the following sections our approaches to creating supramolecular materials by design in which molecules are programmed to

Figure 1. Molecular-graphics representation of targeted supramolecular nanostructures with shapes and sizes defined by the number of molecules in the cluster and the aggregation mode.

42

MRS BULLETIN/APRIL 2000

Self-Assembly of Organic Nano-Objects into Functional Materials

Figure 2. (a) Transmission electron micrograph of stacked two-dimensional polymers with well-defined thickness but irregular x-y dimensions.6 (b) Schematic representation showing how layered molecules of high aspect ratio are connected by covalent bonds within three planes to form two-dimensional macromolecules.

form either zero- or one-dimensional structures. In zero-dimensional structures, x, y, and z dimensions are extremely small and in the range of a few nanometers.

Rodcoil Molecules
The formation of supramolecular aggregates of low dimensionality has been approached by our laboratory using rodcoil molecules. These molecules, which may have oligomeric or polymeric dimensions, are segmented structures in which one block is molecularly rigid and therefore is rodlike in character. One or more other segments share the same covalent backbone with the rigid one, but have low torsional energies (see Figure 3). The blocks with low torsional energies are coil-like because of the large conformational space they could explore as isolated molecules. We and others have previously studied these macromolecules or oligomers both

Figure 3. Schematic representation of three types of diblock rodcoil molecules having rodlike blocks that are either (a) semiflexible, (b) planar, or (c) cylindrical in shape. All three molecules contain a molecularly flexible segment.

experimentally1420 and theoretically.2127 They are interesting as targets for the selfassembly of supramolecular structures with well-defined shapes, since rod segments are likely to encounter nontrivial sterics as they pack within van der Waals distances. For example, they are not likely to pack randomly in the interior of micelles or even radially in the interior of cylinders, thus expanding the shape space one could access as a result of molecular aggregation. The molecularly stiff segments, commonly aromatic in structure, are likely to aggregate or crystallize and therefore reduce the conformational entropy of coil segments. This will set the stage for entropy-controlled finite aggregation of molecules. Furthermore, it is likely that high contact energies between chemical units in rigid versus flexible segments will cause them to nanophaseseparate when aggregation occurs. This nanophase separation will result in rigid sectors and soft disordered sectors within one aggregate and thus help define characteristic shapes and chemical maps. Figure 4 shows examples of possible rodcoil molecules and the aggregates they could form; some are zero-dimensional, and others are one-dimensional structures. The arguments for self-assembly of rodcoil molecules may apply to some extent to biological macromolecules as well. In biology, -helices and -strands are rigid motifs that aggregate in ordered structures connected by more flexible coil-like loops.28 Even though polypeptide-folding rules are clearly complex, rod-to-coil volume ratio must play a key role in the thermodynamics of self-assembly. In synthetic rodcoil molecules of the type we investigate, theoretical work has in fact suggested that relative volumes of rod and coil segments deeply affect their phase behavior.21,22,27 Our laboratory has demonstrated this principle by showing that molecules with high coil-to-rod volume fractions do not form ordered structures by self-assembly.29,30 We have identified through our research at least two fundamentally different mechanisms to mediate the finite aggregation of molecules into supramolecular structures. In one mechanism, twosegment rodcoil molecules form finite structures that are either one-dimensional or zero-dimensional, depending essentially on the molecular length of flexible segments. As explained in the next section, supramolecular size and dimensionality may be controlled by the entropic cost of stretching coil segments upon aggregation of rod segments. This mechanism is likely to be important in rodcoil molecules of polymeric dimension. In

MRS BULLETIN/APRIL 2000

43

Self-Assembly of Organic Nano-Objects into Functional Materials

Figure 5. Fourier filtered transmission electron micrograph image of a hexagonal superlattice formed by aggregates of diblock rodcoil polymers. The core of each disklike structure is formed by the aggregation of rod segments, and the periphery contains splaying coil segments.

Figure 4. Schematic representation of possible rodcoil molecules and the aggregates formed by them. The schematic shows (a) mushroomlike clusters in which rodcoil molecules aggregate parallel to each other and thus form supramolecular entities lacking a center of inversion D h; (b) disks in which rod segments are either parallel or perpendicular to the disk normal, and rods are either bilayer, interdigitated-monolayer, or tilted aggregates; (c) ellipsoidal aggregates with similar rod-segment arrangements as shown for disks; and (d) strips of one-dimensional aggregates that may also lack a center of inversion, in analogy to the mushroom-shaped nanostructures.

another mechanism, precursor molecules with a more complex architecture have the capacity to turn on repulsive forces among segments as aggregation takes place. In this case, the dissipation of repulsive forces is considered to be an important factor in the control of supramolecular size and dimensionality. One mechanism is mediated by coil-unfolding and the other by coil sterics. We explain in the following sections how zerodimensional and one-dimensional structures can form through both mechanisms.

Zero-Dimensional Nanostructures
We reported a few years ago the formation of disklike nanostructures that selforganize into hexagonal superlattices.15 The rodcoil molecules that led to these zero-dimensional structures with dimensions of the order of 10 nm had rather long coil segments, with an average of 6070 structural units of isoprene (the repeat units of natural rubber). The rod segments were all chemically identical, whereas the coil segments had structural diversity both in length and chemical sequence as a result of different modes of

addition among isoprene units.15 Given the structural diversity among coil segments, these molecules cannot form threedimensional crystals. We also did not find any electron-diffraction evidence of localized nanoscale crystallization among the chemically identical rod segments. This may well be the consequence of very long coil segments relative to rod-segment dimensions. A micrograph of these nanostructures is shown in Figure 5. As explained in the next section, shorter coil segments produce elongated, onedimensional structures, but crystallization of rod segments still does not occur. Our interpretation was that longer coils resist unfolding (stretching), and thus aggregation is quenched at lower supramolecular molar mass. This early quenching allows flexible segments to splay around aggregated rod segments, forming zerodimensional, disklike structures. One of our goals is to be able to program designed molecules to form a broad spectrum of cluster shapes, especially shapes of low symmetry. This could include access to nanotubes of any organic structure and their uniaxially aligned

condensed phases, stacked structures of two-dimensional objects, and superlattices of regularly sized objects with defined shapes (see Figure 6). The triblock rodcoil structures shown in Figure 7 were, in fact, found to form nanostructures lacking a center of inversion as a result of parallel aggregation of molecules. Figure 7 shows four different examples of these zero-dimensional clusters, varying both in size and chemical composition, and all containing nanocrystalline and amorphous sectors. In order of increasing size, the smallest nanostructures contain fluorinated surfaces;31 the next largest ones have a photoluminescent sector containing phenylene vinylene segments,30 followed by others with a highly adhesive base of phenolic groups;32 and the largest ones have a diode architecture with amorphous compartments containing triphenylamine hole-accepting groups and oligo(phenylene vinylene) crystalline domains that could accept electrons.33 There might be a correlation between size and polydispersity of the nanostructures, but this remains an open question. The common structural feature of the molecules shown in Figure 7 is a coil segment with a larger cross-sectional area than other blocks of the molecule. We therefore believe that steric forces among the wider segments play a role in their formation. In fact, for such short coil lengths, we would expect layered twodimensional structures to form. This statement is based on our earlier work on two-dimensional polymers and on the

44

MRS BULLETIN/APRIL 2000

Self-Assembly of Organic Nano-Objects into Functional Materials

Figure 6. Molecular-graphics representation of ordering among supramolecular nanostructures of various shapes. (a) Platelike two-dimensional structures are expected to stack in condensed phases; (b) cylinders would align uniaxially; and (c), (d) zero-dimensional nanostructures with regular sizes and shapes could form superlattices with symmetries that may differ from those prevailing within the crystalline sectors of the nanostructures. Reprinted with permission from Science 276 (5311) p. 384. Copyright 1997, American Association for the Advancement of Science.

general structure of molecules known to form smectic phases.34 Also, in this case, identical rod segments do crystallize, a critical event in the energetics of nanostructure formation. Rod-segment crystallization here could be partly related to the short length of the coil segments. The absence of inversion symmetry in these structures makes them analogous to many important nano-objects in biology. Many proteins have similar shapes, particularly the highly functional transmembrane structures that transmit signals from an external environment to the interior of the cell. Such structures tend to have chemically different sectors that either prefer the hydrophobic interior of cell membranes or the hydrophilic external environment. Their positions allow them to mediate biological events through ligand-receptor interactions. In analogy to such proteins, the nanostructures of Figure 7 have distinct chemical sectors defined by the aggregation mode and the segmented nature of the molecules. The thermodynamics behind formation of finite aggregates of molecules in parallel configuration is not entirely clear. The

larger cross section of one terminal segment should introduce curvature in the embryonic stages of aggregation. This curvature dissipates repulsive forces experienced by the widest segments as molecules come into van der Waals contact. We assume the resultant splaying of coil segments eventually defines the cluster. Using molecular dynamics, we have established that a higher aggregation number leads to a decrease in the number of conformations accessible to splaying segments, as they dissipate the repulsive forces turned on by self-assembly.35 Thus the entropic cost to coil segments covalently grafted to nanocrystals could be an important factor leading to finite aggregation of these molecules. It is intriguing to explore why these triblock rodcoil molecules tend not to interdigitate and form centrosymmetric nanostructures. It is possible that in the absence of a strong dipole moment parallel to their backbone, their characteristic mushroom geometry optimizes contact among all segments in parallel aggregation. This could be the basis of their ferrophilic as opposed to antiferrophilic preferences during self-assembly.

One of the interesting aspects of materials created by these clusters lacking a center of inversion is their tendency to form films that reveal evidence of a macroscopically polar structure. At the same time, transmission electron microscopy of ultra-microtomed slices of these films, as well as x-ray diffraction, shows that these films self-assemble into layered structures. The polar nature of the films is expressed in both their ability to exhibit second-harmonic generation32,36 and piezoelectricity.37 Figure 8 shows a measurement of spontaneous polarization observed in unpoled films composed of clusters (after annealing) and the corresponding piezoelectric response of the same film. Given these observations, one infers that net polar stacking of nanostructures occurs in the volume sampled by these experiments. Most likely the structure prevailing in these materials, based on the magnitude of measured optical and electrical coefficients, is a polydomain structure with low overall polarization.37 It is not clear at this point if such polydomain structures have kinetic or thermodynamic origin. Nonetheless, when some of these systems are exposed to external electric fields, a very high degree of alignment is achieved, suggesting a cooperative alignment of nanostructures and their domains. This has been suggested by experiments on piezoelectric activity in these films.37 In the absence of the field, we believe substrates must play a role in establishing orientation and polarity on macroscopic scales.

Biological Analogies
Biology has many examples of proteins and mesoscale aggregates of proteins lacking a center of inversion. The polar nature of these structures has a key role to play in biological function. Figure 9 illustrates some of these structures. Most important to this article is the structure of -hemolysin, which is a seven-protein aggregate with a non-centrosymmetric mushroom shape.38 The aggressive human pathogen Staphylococcus aureus uses the asymmetric nature of this object to implant its stem into the hydrophobic compartment of cell membranes and the hydrophilic nature of its cap to stabilize it in the extracellular space. Its fascinating function relies on the existence of a rather large pore of about 16 , compared with biological ion channels,39 running parallel to its long axis. It is through this pore that RNA macromolecules of the pathogen can invade human cells. Having learned how to achieve the parallel self-assembly of molecules in large quantities, one could easily consider many technological con-

MRS BULLETIN/APRIL 2000

45

Self-Assembly of Organic Nano-Objects into Functional Materials

Figure 7. Transmission electron micrographs of four different supramolecular materials composed of nanostructures (right column). The contrast in the micrographs is mainly from diffraction due to the presence of crystalline compartments in all clusters. At left, molecular-graphics representations are shown of the supramolecular nanostructures and the corresponding triblock molecules that form them. From (a) to (d), the average number of molecules per cluster N increases from 23 to 144 molecules. Note differences in polydispersity of the clusters in the four different materials and the formation of two-dimensional superlattices in (c) and (d). The cluster in (a) contains a fluorinated base, the one in (b) contains a photoluminescent sector, the one in (c) contains a base of very adhesive phenolic groups, and the largest cluster (d) has the supramolecular architecture of a diode, with a p-type hole-transporting sector of triphenylamine residues at the top and a stem containing an n-type emissive sector of oligo(phenylene vinylene) segments. Note that the measured bar in (d) is different from that in (a) through (c).

Figure 8. (a) A strongly photoluminescent solution of a triblock molecule and the corresponding supramolecular solid film obtained after solvent evaporation (both samples are being excited with 365-nm radiation and emit blue light). (b) Increasing values of polarization measured in the solid film shown in (a) as a function of temperature and the value measured after the film is cooled to room temperature (after annealing). (c) Changes in film dimension h with applied voltage, revealing a piezoelectric response in the self-assembled supramolecular film. E and R are the electric-field intensity and correlation coefficient, respectively. Reprinted with permission from Angew. Chem. 39 (2000) in press. Copyright 2000, Wiley-VCH Verlag GmbH.

46

MRS BULLETIN/APRIL 2000

Self-Assembly of Organic Nano-Objects into Functional Materials

Figure 9. (top left) Mushroom-shaped aggregate of seven polypeptides known as -hemolysin and produced by a human pathogen.38 (bottom left) Molecular-graphics representation of a supramolecular cluster observed experimentally with similar dimensions to -hemolysin (see Figure 7). (top right) Schematic representation of the polar stacking of protein dimers in tubulin, present in the cytoskeleton of cells.40 (bottom right) Schematic representation of polar stacking of layered nanostructures observed in a supramolecular material investigated in our laboratory.

cepts related to molecular transport into and out of materials mediated by nanostructures implanted on surfaces. This will require, of course, developing strategies to core the assemblies. For protein aggregates, the polar structure of tubulin formed by the aggregation of - dimers has a key role to play in vectorial transport inside cells.40 The polar structure of actin filaments must also be important in cytoskeleton dynamics.41 Polar stacking of proteins is analogous to the formation of polar aggregates by zero-dimensional nanostructures lacking a center of inversion, as observed in our experiments.

One-Dimensional Assemblies Strip Phases


An important direction in supramolecular materials science is to create general synthetic strategies for the creation of onedimensional structures. The well-known carbon nanotubes belong to this subset of molecular objects, but we need much more general strategies that could accept a broad range of organic and inorganic chemical structures. One-dimensional architectures could be very useful as shuttling conduits for ions or molecules in microfluidics or between cells in biomedi-

cal sensors or devices. Our laboratory has developed three distinct strategies for creating one-dimensional structures. One of them is the self-organization of rodcoil molecules into strip phases in which rod segments aggregate to produce elongated structures with aspect ratios in the range of 10. These phases appear to form at characteristic rod-to-coil volume ratios in diblock rodcoils,16 presumably based on the entropic cost associated with the unfolding (stretching) of the coil segments. We have also observed them in the triblock systems, which we believe are mediated by steric forces.30 In this case, onedimensionality could have its origin in the very strong tendency of specific rod segments to aggregate and crystallize. The one-dimensional strip has a much higher aggregation number relative to the zerodimensional clusters (by an order of magnitude). Micrographs of these phases are shown in Figure 10.

Figure 10. Transmission electron micrographs of one-dimensional structures formed by the aggregation of (a) diblock and (b) triblock rodcoil molecules (strip phases). The inset in (a) is a Fourier filtered image; point a represents a dislocation and point b represents a disclination of type s = +1/2.

Cylinders and Ribbons


We have also developed a templating approach to one-dimensional structures in which polyunsaturated and therefore polymerizable amphiphiles are confined at the interface of hydrophobic and hydro-

philic compartments of hexagonal phases formed by surfactants. These systems of polyfunctional, reactive amphiphiles within amphiphilic templates have the capacity to form one-dimensional, cylindrical polymers with molar masses on the order of 42 106 Da.42 Furthermore, these cylindrical macromolecules templated by the hexagonal mesophase form microscopic droplets similar to those observed in nematic liquid crystals, with a high orientational order parameter. More recently, we have discovered systems in which ribbonlike supramolecular structures selfassemble from solution.43 The ribbon is a bimolecular string formed by hydrogen bonds with a width of approximately 10 nm. Thus the structure has some architectural resemblance to DNA strands. These supramolecular ribbons form in organic solvents and have the capacity to transform these solutions into birefringent gels at exceedingly low concentrations, of the order of 0.3 wt%. Aremarkable observation made

MRS BULLETIN/APRIL 2000

47

Self-Assembly of Organic Nano-Objects into Functional Materials

recently is that styrene gels containing the one-dimensional structures polymerize to form birefringent solids and melts that are essentially 99% polystyrene by weight, a classical organic glass. We believe the mechanisms involved in the formation of these structures are rather complex, involving hydrogen-bonding, sterics, and precursor-solvent interactions as well.

Conclusions
The use of supramolecular chemistry to design highly functional materials is a wide open field for materials-science research. Our ability to design structures of low dimensionality will introduce novel properties in designed materials that are extremely useful in biological systems. Two-dimensionality of protein networks is a great tool for controlling the ability of cells to change shape reversibly.44 In the world of nanotechnology, we may benefit enormously from mesoscale materials that have such capabilities as a result of external stimuli. Descending to onedimensional structures, we might be able to mediate transport phenomena and control the orientation of highly anisotropic mesoscopic objects. In this regard, biology shows us the power of cytoskeleton dynamics in

cells with one-dimensional polar structures such as actin fibers and tubulin. In our opinion, absolute control over the chemistry, shape, and size of zero-dimensional structures should be the ultimate goal. This could lead to a toolbox for materials of truly synthetic, proteinlike objects that differ greatly in chemical structure from those observed in biology. The achievement of this goal will not only yield an enormous array of superlattice materials with the same richness as organic crystals, but also the opportunity to mimic, in synthetic systems, the mysteries of signal transduction in biology. One exciting possibility would be to bring ligandlike and receptorlike nanostructures into contact to trigger changes in the physical properties of materials or in their shapes or dimensions.

Acknowledgments
Our laboratorys work on supramolecular materials has been supported by the U.S. Department of Energy, the U.S. National Science Foundation, the U.S. Office of Naval Research, and the U.S. Army Research Office.

References
1. J.-M. Lehn, Supramolecular Chemistry (VCH, New York, 1995). 2. D. Philip and J.F. Stoddart, Angew. Chem., Int. Ed. Engl. 35 (1996) p. 1154. 3. J.L. Atwood, J.E.D. Davies, D.D. MacNicol, and F. Vgtle, Comprehensive Supramolecular Chemistry (Pergamon Press, New York, 1996). 4. P.-G. de Gennes, The Physics of Liquid Crystals (Oxford University Press, New York, 1975). 5. S.I. Stupp, S. Son, H.C. Lin, and L.S. Li, Science 259 (1993) p. 59. 6. S.I. Stupp, S. Son, L.S. Li, H.C. Lin, and M. Keser, J. Am. Chem. Soc. 117 (1995) p. 5212. 7. S. Mori and M.J. Wadati, J. Phys. Soc. Jpn. 62 (1993) p. 3864. 8. F.F. Abraham and D.R. Nelson, Science 249 (1990) p. 393. 9. F.F. Abraham and D.R. Nelson, J. Phys. (Paris) 51 (1990) p. 2653. 10. F.F. Abraham and M. Kardar, Science 252 (1991) p. 419. 11. Y. Kantor and K. Kremer, Phys. Rev. E 48 (1993) p. 2490. 12. D.C. Morse, I.B. Petsche, G.S. Grest, and T.C. Lubensky, Phys. Rev. A 46 (1992) p. 6745. 13. V. LeBonheur, PhD thesis, University of Illinois at Urbana-Champaign, 1996, p. 109. 14. L.H. Radzilowski, J.L. Wu, and S.I. Stupp, Macromolecules 26 (1993) p. 879. 15. L.H. Radzilowski and S.I. Stupp, Macromolecules 27 (1994) p. 7747. 16. L.H. Radzilowski, B.O. Carragher, and S.I. Stupp, Macromolecules 30 (1997) p. 2210.

17. A. Douy and B. Gallot, Polymer 28 (1987) p. 147. 18. W. Li, H. Wang, L. Yu, T.L. Morkved, and H.M. Jaeger, Macromolecules 32 (1999) p. 3034. 19. R.S. Saunders, R.E. Cohen, and R.R. Schrock, Macromolecules 24 (1991) p. 5599. 20. D. Marsitzky, T. Brand, Y. Geerts, M. Klapper, and K. Mullen, Macromol. Rapid Commun. 19 (1998) p. 385. 21. A. Halperin, Macromolecules 23 (1990) p. 2724. 22. E. Raphal and P.-G. de Gennes, Makromol. Chem., Macromol. Symp. 62 (1992) p. 1. 23. K.S. Schweizer, Macromolecules 26 (1993) p. 6050. 24. A.N. Semenov and S.V. Vasilenko, Sov. Phys. JETP (Engl. Transl.) 63 (1986) p. 70. 25. A.N. Semenov, Mol. Cryst. Liq. Cryst. 209 (1991) p. 191. 26. J.D. Vavasour and M.D. Whitmore, Macromolecules 26 (1993) p. 7070. 27. D.R. Williams and G.H. Fredrickson, Macromolecules 25 (1992) p. 3561. 28. W.F. DeGrado, C.M. Summa, V. Pavone, F. Nastri, and A. Lombardi, Annu. Rev. Biochem. 68 (1999) p. 779. 29. M.U. Pralle, C.M. Whitaker, P.V. Braun, and S.I. Stupp, Macromolecules in press. 30. G.N. Tew, M.U. Pralle, and S.I. Stupp, J. Am. Chem. Soc. 121 (1999) p. 9852. 31. E.R. Zubarev, M.U. Pralle, L.M. Li, and S.I. Stupp, Science 283 (1999) p. 523. 32. S.I. Stupp, V. LeBonheur, K. Walker, L.S. Li, K. Huggins, M. Keser, and A. Amstutz, Science 276 (1997) p. 384. 33. G.N. Tew, M.U. Pralle, and S.I. Stupp, Angew. Chem., Int. Ed. Engl. 39 (2000) p. 517. 34. G.W. Gray and J.W.G. Goodby, Smectic Liquid CrystalsTextures and Structures (Heyden and Sons, Philadelphia, 1984). 35. M. Sayar and S.I. Stupp, Self-Assembly of Rodcoil Molecules into Nanostructures, presented at Materials Research Society Meeting, Boston, November 1999. 36. G.N. Tew, L.M. Li, and S.I. Stupp, J. Am. Chem. Soc. 120 (1998) p. 5601. 37. M.U. Pralle, K. Urayama, G.N. Tew, D. Neher, G. Wegner, and S.I. Stupp, Angew. Chem., Int. Ed. Engl. in press. 38. L. Song, M.R. Hobaugh, C. Shustak, S. Cheley, H. Bayley, and J.E. Gouaux, Science 274 (1996) p. 1859. 39. H. Lodish, D. Baltimore, A. Berk, S.L. Zipursky, P. Matsudaira, and J. Darnell, Molecular Cell Biology (Scientific American, New York, 1995) p. 943. 40. E. Pennisi, Science 279 (1998) p. 176. 41. P.W. Hochachka, Proc. Natl. Acad. Sci. U.S.A. 96 (1999) p. 12233. 42. S. Eftekharzadeh and S.I. Stupp, submitted for publication, 1999. 43. E.R. Zubarev, M.U. Pralle, E. Sone, and S.I. Stupp, submitted for publication, 1999. 44. A. Elgsaeter, B.T. Stokke, A. Mikkelsen, and D. Branton, Science 234 (1986) p. 1217. s

FEDERATION OF MATERIALS (FMS) DIGITAL REVOLUTION AND MATERIALS COMMUNITY


May 21-24, 2000 Washington, DC www.materialssocieties.org

48

MRS BULLETIN/APRIL 2000

Вам также может понравиться