You are on page 1of 16

Self-Assembled Lipid Nanotube Hosts: The Dimension Control for Encapsulation of Nanometer-Scale Guest Substances

TOSHIMI SHIMIZU Nanoarchitectonics Research Center (NARC), National Institute of Advanced Industrial Science and Technology (AIST), Tsukuba Central 5, 1-1-1 Higashi, Tsukuba, Ibaraki 305-8565, Japan

Received 30 September 2005; accepted 1 December 2005 DOI: 10.1002/pola.21619 Published online in Wiley InterScience (www.interscience.wiley.com).

Supramolecular nanotube hosts with precisely controlled inner or outer diameters have been synthesized by self-assembly of unsymmetrical bolaamphiphilic monomers or glucopyranosylamide lipids, respectively. Time-resolved uorescent measurement using 8-anilinonaphthalene-1-sulfonate (ANS) as a probe revealed that the water conned in a cardanyl-b-D-glucopyranoside lipid nanotube has relatively lower solvent polarity corresponding to that of propanol than bulk water. Extensively developed hydrogen bond networks also characterize the conned water in comparison to the case in bulk water. Encapsulation ability of the glucopyranosylamide lipid nanotube has been examined by lling the lyophilized LNTs with gold or silver nanoparticles, ferritin, or magnetic crystals. Filling the unsymmetrical bolaamphiphile nanotube possessing positively charged inner surfaces with negatively charged polymer beads or ferritin proved to be successful without depending C on capillary action. V 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44:
ABSTRACT: 51375152, 2006

Keywords: amphiphiles; nanocomposites; nanotechnology; self-assembly; supramolecular structures

INTRODUCTION
Recently a number of amphiphilic molecules programmed with desired functionalities have been known to self-assemble in various liquid media to give tubular structures with nanometer-scale dimensions and high-axial ratios.13 These ndings not only stimulate intriguing molecular design of tube-forming compounds but also inspire novel methodology for the formation of organic nanotubes.26 A variety of applications to use such discrete, organic nanotubes as nanochannels7 and
Correspondence to: T. Shimizu (E-mail: tshmz-shimizu@ aist.go.jp)
Journal of Polymer Science: Part A: Polymer Chemistry, Vol. 44, 51375152 (2006)
C V 2006 Wiley Periodicals, Inc.

nanotemplates8,9 are now emerging in nanobiotechnology and new materials science elds. The rst discovery of noncovalent synthesis of nanotubes from simple amphiphilic compounds goes back to the concurrent works by three research groups in Japan and USA.1012 At that time, there had been no terminology of nanotube since Iijima was the rst to dene this word for tubular carbon nanomaterials.13,14 Organic tubular architectures can be categorized into several groups including hollow helices, their aggregates, stacked rings, stacked rosettes, and molecular assemblies, as dened by Hecht and coworkers (Fig. 1).2 Among them, self-assembled nanotube structures have attracted much attention due to relatively more facile synthesis of tube-forming amphiphiles and the large-scale production of resultant nano5137

5138

SHIMIZU

Figure 1. A variety of tubular architectures with open ends, produced by self-assembly of a molecular building block: (a) aggregated helices from a hollow helix, (b) stacked rings from a ring, (c) stacked rosettes from a rosette, and (d) monolayerbased LNT from an unsymmetrical bolaamphiphile. [Color gure can be viewed in the online issue, which is available at www.interscience.wiley.com.]

tubes.3 They are also of great interest in terms of their new properties and resemblance to wellknown carbon nanotubes. When focusing on noncovalent synthesis of the nanotube structures, there are several practical methods to yield the nanotubes.3 Those noncovalent approaches involve, for example, chiral molecular self-assembly,1517 packing-directed self-assembly,18,19 polymer assembly,2022 molecular sculpting,2326 and template-directed synthesis using a porous template.2730 In this short review, we focus on our recent results on the molecular self-assembly of sugarbased amphiphiles into well-dened nanotube architectures. When dening the terminology of tubular structures, one can add a species name as a prex: for example, carbon nanotube, silica nanotube, or gold nanotube. To confer generality on the terminology of nanotubes, we use the term lipid nanotube (LNT) hereafter. We also describe the properties and function associated with the nanometer-scale hollow cylinder provided by the supramolecular nanotubes. Initially, we start to discuss the polarity and structures of water conned in the nanometer-sized geometries. The resultant hollow cylinder structures of the LNT with

nanometer-scale dimensions are also of great interest as tubular host nanomaterials that can encapsulate biomacromolecules or nanometer-scale guest objects. We then describe the recent progress in the dimension control of nanotube hosts based on amphiphilic molecules. However, selective lling of the inner hollow of the LNTs with foreign nanomaterials more than 10 nm wide has never been addressed. Here we highlight onedimensional (1D) organization of gold nanocrystals or spherical proteins in a conned nanospace of the uniform glycolipid nanotubes from a glucopyranosylamide lipid. It should be noted here that corresponding guest substances are more than 10 times larger in dimension in comparison to those used for conventional host-guest chemistry31,32 using cyclodextrin and other macrocyclic or spherical host molecules.

PROPERTIES OF WATER CONFINED IN THE HOLLOW CYLINDER OF A GLYCOLIPID NANOTUBE


Water conned in nanometer-sized restricted geometries often exhibit different and sometimes
Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

SELF-ASSEMBLED LIPID NANOTUBE HOSTS

5139

Figure 2. TEM image of the glycolipid nanotube self-assembled from 1 and schematic illustration of 3D model for the molecular packing in the LNT. The molecules 1 form an interdigitated bilayer structure to produce three bilayer sheets as a membrane wall 1015 nm thick. [Color gure can be viewed in the online issue, which is available at www.interscience. wiley.com.]

peak of the bulk water (kmax 525 nm)54 or the water pool in the sodium bis(2-ethylhexyl)sulfosuccinate (AOT) reversed micelles (kmax 490 nm), the corresponding peak of the conned ANS remarkably shifted to kmax 467 nm at a steady state (Fig. 3).53 This nding means that the local solvent polarity [ET(30)] of the conned water decreased to 50 kcal/mol, whose value really corresponds to that of propanol 20% lower than that in the bulk water [ET(30) 63 kcal/mol].55 On the other hand, attenuated total reectance infrared measurement revealed that the OH stretching band for the conned water consists of several components attributable to extensively developed hydrogen bond networks, generally observed at low temperatures.53 Abundant sugar OH groups covering the inner surfaces of the glycolipid nanotube from 1 are responsible for the immobilization

unexpected characteristics in comparison to bulk water. Various advanced spectroscopic techniques and computer simulations have unveiled the properties of water conned in micelles and microemulsions,3338 nanometer lms,3941 nanoporous silica,42 and other kinds of conned structures. In particular, cylindrical hollows of the LNTs with high-axial ratios have many potential applications as nanochannels for nanouidic devices,4346 templates for metal nanowire formation,4749 and size-selective pores for analytical or storage devices.50,51 Self-assembly in water of cardanol-based glycolipids 1 allowed us to obtain the LNTs used for the next loading experiment. The resultant LNTs have inner diameters of 1015 nm and outer diameters of 4050 nm (Fig. 2).52 To load a uorescent probe molecule, 8-anilinonaphthalene-1-sulfonate (ANS) selectively into the hollow cylinder of the LNT from 1, we lled lyophilized LNTs with an ANS solution (c 1 104 M) by capillary action and washed the ANS solutions remaining on the outside of the LNT with an excess amount of water thoroughly.47,53 Time-resolved uorescent measurement of the ANS revealed the structural properties of the conned water in the LNT hollow cylinder. As compared with the uorescent
Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

Figure 3. Fluorescence spectra of ANS (a) in bulk water, (b) in the water pool in sodium AOT reversed micelles, and (c) in the conned water in the hollow cylinder of the self-assembled LNT from 1.

5140

SHIMIZU

and structuring of the water molecules in the region adjacent to the inner surfaces.

Figure 4. Schematic illustration of a LNT consisting of MLMs of unsymmetrical bolaamphiphile 2(n) (n 14, 16, 18, and 20). The inner and outer surfaces of the LNT are covered with carboxyl groups and glucose moieties, respectively. [Color gure can be viewed in the online issue, which is available at www.interscience. wiley.com.]

CONTROL OF INNER DIAMETERS FOR UNSYMMETRICAL BOLAAMPHIPHILE NANOTUBES


Unique properties of the conned water in the glycolipid nanotube of 1 suggest that the liquid nanospace shaped by the LNT cylindrical hollow functions as a stable nanocontainer for biomacromolecules such as proteins and DNA. Such nanospace should also be favorable to selective encapsulation and controlled release of the macromolecules. For this purpose, structural optimization of the nanotube dimension including inner and outer diameters, wall thickness, and length is of great importance for a given application. To aim at lling the LNT hollow cylinder with 1050 nm scaled macromolecules or nanomaterials, we have to freely control the inner diameters of LNTs at rst and then need to design rationalized functionalities of their inner surfaces. The inner diameters of the hollow are sensitively correlated with the connement behavior of guest substances encapsulated.48 However, nobody has ever achieved the regulation of inner diameters for the self-assembled LNTs with cylindrical hollows more than 10 nm wide.3 Very recently, we have succeeded in the control of inner diameters in 1.5-nm steps as well as the decoration of inner and outer surfaces with unsymmetrical functionalities (Fig. 4).18 By

making use of the asymmetry of heteroditopic 1,x-amphiphiles having two hydrophilic headgroups differing in size or properties, we synthesized novel unsymmetrical bolaamphiphiles 2(n) to form unsymmetrical monolayer lipid membranes (MLMs). Fuhrhop et al. were the rst to propose this strategic concept.56,57 However, except for limited examples of vesicular or brous assemblies,58,59 and crystal structures,60 most attempts with unsymmetrical bolaamphiphiles have resulted in the formation of symmetrical MLMs.61,62 Otherwise, molecular orientation in MLMs has been random or unknown. We recently found that the relationship between the actual membrane stacking periodicity obtained from X-ray diffraction measurement and the extended molecular length of 2(n) can give a good clue to assign the polymorph and polytype of the MLM arrangement within the tubular assemblies.60,63 Actually, by comparing the stacking periodicity of each dehydrated nanotube of 2(n) (n 12, 14, 16, 18, or 20) (Fig. 5) with its molecular length, we were able to dene the polymorphs and polytypes of the MLM. As a result, except for 2(12) the obtained nanotubes from 2(n) (n 14, 16, 18, and 20) have been evidenced, for the rst time, to consist of unsymmetrical MLMs, in which the bolaamphiphile molecules pack in a parallel fashion.18 Thus, the innermost surfaces of the nanotubes from 2(n) (n 14, 16, 18, and 20) are covered with carboxy headgroups and the outer surfaces with 1Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

SELF-ASSEMBLED LIPID NANOTUBE HOSTS

5141

Figure 6. TEM image of the LNT self-assembled from 1(20). The sample was negatively stained with phosphotungstate.

Figure 5. XRD patterns in the small-angle region for the dehydrated LNTs from 2(n) (n 12, 14, 16, 18, and 20).

glucosamide headgroups. Moreover, changing the spacer carbon numbers from n 14 to n 20 enabled us to produce LNTs with different average inner diameters 17.7, 18.7, 20.8, and 22.2 nm by self-assembly (Fig. 6). Table 1 shows that the calculated inner diameters for the resultant LNTs are well consistent with the observed ones by transmission electron microscopy (TEM).

micrometer-scale space. Doping of short-chain saturated homologues 3(m) to the self-assembly of diacetylenic phospholipids 4(m,n) proved to inuence the outer diameters of the resultant nanotubes.64 Subtle change in the molecular structure of 4(8,9) also inuence the outer diameters of the resultant LNTs.65,66 We have recently found that the numbers and incorporation position of a cis-double bond in the hydrophobic chains of glycolipids drastically affect the self-assembly behavior into tubular morphologies. In particular, the efciency and homogeneity of the tube formation, and the size distribution of outer diameters are sensitive to the unsaturation motifs.6769 We examined the selfassembly behavior of a series of long-chain 4alkanoylamino-phenyl glucosides 5, 6, 7, and 8 that differ in number of a cis-double bond in the hydrophobic part. With the increase in the

Table 1. Inner Diameters of Lipid Nanotubes Consisting of Unsymmetrical Bolaamphiphiles 2(n), Based on Calculations and TEM Observations Inner Diameters Observed by TEM (Dobs-in, nm) 20.6 17.7 18.7 20.8 22.2 6 6 6 6 6 1.9 1.6 1.6 2.3 2.1

Chain Length of 2(n) n n n n n 12 14 16 18 20

Inner Diameters Calculateda (Dcalc-in, nm) 15.6 17.1 18.6 20.2 21.7

CONTROL OF OUTER DIAMETERS FOR GLUCOPYRANOSYLAMIDE LNT


The manipulation of outer diameters for LNTs, which are generally in the range of 401000 nm, directs the handling ability of the LNTs as well as the number of integration in a dened
Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

a If a nanotube consists of unsymmetrical MLMs in which relatively larger headgroups occupy the outer surface of the MLM, the innermost diameter (Dcalc-in) can be dened by the following equation: Dcalc-in 2asL/ (al as), where as and al are the cross-section area of the small and large headgroup, respectively, and L is the molecular length. The way to induce this equation is reported in the supporting information of ref. 18.

5142

SHIMIZU

degree of unsaturation, the glycolipid tends to produce homogeneous tubular assemblies.67

packing within solid bilayer membranes is known to play a key role in shaping twisted, coiled, or tubular structures.16,17,7072 Therefore, we carried out CD measurement for the aqueous solutions containing each tubular structure to conrm the order of chiral molecular packing. The intensity of the observed CD band about 235 nm increased in the order of 11, 9, and 10 (Fig. 8). The uniformity of the outer diameters for the resultant LNTs also increased in agreement with this tendency. Continuum theory for the LNT formation predicts that the magnitude of molecular chirality and the molecular tilting with respect to the bilayer planes are responsible for the outer diameters of the resultant nanotubes.3,7376 A bent conformation of the 11-cis derivative 10 will induce the most effective chiral packing in the solid bilayer membranes, resulting in the narrowest distribution of outer diameters.

To further optimize the molecular structure for tube-forming compounds, we newly designed a series of glucopyranosylamide lipids 914.68 The location of a cis-double bond in the C18 hydrocarbon chains proved to direct the uniformity and yields of the resultant self-assembled nanotubes. As a result of detailed analysis for the nanotube dimensions, among six homologues the 11-cis derivative 10 self-assembled in water to form uniform nanotube structures with the narrowest distribution of outer diameters in more than 98% yields (Fig. 7 and Table 2).68 Chiral molecular

ENCAPSULATION OF 1050-nm-SCALED GUEST SUBSTANCES INTO LNT


The dimensions of nanoscale biomacromolecules and related biological objects are compatible with the inner diameter sizes of LNTs. Figure 9 shows the typical values of inner diameters available for certain LNTs and compares those dimensions with natural or synthetic, spherical molecular
Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

SELF-ASSEMBLED LIPID NANOTUBE HOSTS

5143

which accommodates ferrihydrite core inside.77 The dimensions of the outer shell and inner core of ferritin are 12 and 7 nm, respectively. Wellknown dendritic polymers, dendrimers, are in the range of 520 nm in diameters.7881 Norwalk-like virus 27 nm wide is one of the smallest spherical viruses. Magnetosome is intracellular compartment in magnetotactic bacterium, which houses highly ordered magnetite crystals.82,83 The spanned sizes of magnetosome are in the range of 50 100 nm. Thus, one may realize that these spherical objects 12100 nm wide should be suitable guests for the self-assembled LNT hosts. Next we describe the successful examples of encapsulation by the optimized or functionalized LNT hosts from 10 or others.

Gold and Silver Nanoparticles By using the morphologically uniform LNTs self-assembled from 10 as nanotube hosts,68 we have for the rst time achieved the 1D conned organization of gold nanocrystals in the LNT hollow cylinder at room temperature.47,48 Figure 10 shows a series of experimental procedures to conne gold nanocrystals exclusively in the hollow. First, to remove the water volume inside the hollow cylinder, we lyophilized the selfassembled LNTs in a vacuum, without damaging the tubular morphologies. Addition of the dried LNT powders into 20% ethanolic aqueous solutions containing hydrogen tetrachloroaurate (HAuCl4) allowed for the lling of the solutions into the vacant LNT hollow by capillary action. The lled LNTs were separated by ltration and were washed thoroughly with water to remove the HAuCl4 species existing on the outer surfaces of the LNT. After redispersing the lled LNTs again in water, photochemical reduction of

Figure 7. FE-SEM images of the LNTs selfassembled from the glucopyranosylamide lipid 10. Scale bar 1000 nm (top) and 100 nm (bottom).

objects having nanoscale dimensions. Although the LNTs obtained so far are reported to give the inner diameters in the range of 101000 nm, most of the inner diameters fall in the range of 10 200 nm.3 For example, an iron-storage protein ferritin is one of the largest spherical proteins,

Table 2. Dimensions of the Self-Assembled Lipid Nanotubes from 9, 10, or 11a Outer Diameter (Dout) [a.v. 6 s.d. (nm)] 214 6 55 200 6 23 220 6 71

Glycolipid 9 10 11

Inner Diameter (Din) [a.v. 6 s.d. (nm)] 76 6 26 61 6 19 73 6 22

Wall Thicknessb [a.v. 6 s.d. (nm)] 60 6 24 54 6 12 63 6 24

a.v., Average value; s.d., standard deviation value. a Evaluated for 250 pieces of lipid nanotubes. b Directly measured the distances of lipid nanotube walls.

Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

5144

SHIMIZU

Figure 8. CD spectra for the molecular self-assemblies from (a) 13, (b) 11, (c) 9, and (d) 10.

the HAuCl4 species in the hollow resulted in the conned production of gold nanocrystals in a \ship-in-bottle" fashion. The use of nonlyophilized LNTs gave very low yields of the lled LNTs with gold nanocrystals (<3%), which contrast to the relatively higher yields ($30%) when using the lyophilized LNTs.48 The TEM image revealed that gold nanocrystals of 310 nm are organized in the conned nanospace of the LNT hollow cylinder (Fig. 10, bottom). Electron diffraction pattern for the selected area supports the exact presence of gold nanocrystals characterized by a fcc phase. However, the aforementioned methodology is only applicable to metal species that are labile to photoreduction in solution systems. Furthermore, there are no efcient ways to regulate the size of metal nanocrystals conned during a series of procedures. On the other hand, metal colloid nanoparticles possess both well-dened shapes and controllable size dimensions, and are easily available for a variety of metals.

Therefore, we attempted to ll the same vacant LNTs from the glycolipid 10 with water-soluble gold or silver nanoparticles using capillary action. Consequently, the glycolipid nanotubes proved to encapsulate the gold or silver nanoparticles in $30% yields.48 When using different size of gold nanoparticles as guests, we found remarkable size effect of the gold nanoparticles on the 1D organization prole in the LNT hollow cylinder. For example, gold nanoparticles 1520 nm wide can align side-by-side along the long axis of the LNT hollow 30 50 nm wide, whereas relatively smaller nanoparticles 13 nm wide ll the LNT hollow with close packing (Fig. 11). However, we failed in loading the LNT with relatively larger gold nanoparticles of more than 50 nm. If metal nanocrystals can be aligned rationally with a desired size, shape, and morphology, one can fabricate metal nanowires by making use of cylindrical hollowness of the LNT [Figs. 10(d,e)]. This methodology has also the advantage that the width of the nanowire can be tuned by the inner diameters size of the LNT. Rational molecular design enables one to produce the LNTs with desired inner diameters.18,19 We realized that complete removal of the LNT organic shell from the LNT-gold nanocomposite by ring in air can allow gold or silver nanoparticles to remain with being conned and eventually results into 1D gold nanowires.48 Thermogravimetric analysis (TGA) revealed that the decomposition of the LNT template from 10 starts at 293.6 8C and burned out at 657.5 8C (Fig. 12). By optimizing the burning temperature between 450 and 1000 8C, we found that a continuous gold nanowire 2030 nm wide was successfully obtained under the ring conditions of T 550750 8C (Fig. 13). However, the formation of silver nanowires was unsuccessful in the similar manner as described earlier. On the other hand, the nanotube self-assembled from the Alzheimers b-amyloid diphenylalanine was demonstrated to serve as molds for casting silver nanowires 20 nm wide.50 Matsui et al. have also developed magnetic nanotube fabrication by lling the peptide nanotube, developed by our own research group,84,85 with bacterial magnetosome.49 Ferritin Ferritin has a spherical protein shell 12 nm wide, composed of 24 protein subunits, the inner
Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

SELF-ASSEMBLED LIPID NANOTUBE HOSTS

5145

Figure 9. Comparison of the dimensions for typical LNTs and natural or synthetic, spherical molecular objects. The images of ferritin and dendrimer were reproduced with permission from Dr. Ichiro Yamashita (Matsushita Electric Industrial) and Prof. William A. Goddard III (Materials and Process Simulation Center, California Institute of Technology). The image of Norwalk-like virus was reproduced with permission from P. Bijkerk, Infectieziekten Bulletin, 2002, 13, 271-272, Copyright 2002 RIVM/ CIE.

cavity of which accommodates ferrihydrite core in vivo.77 Therefore, 1D arrangement of ferritin is expected to provide nanodots or quantum dots with 12-nm intervals as a key component of single-electron devices.8688 Manns research group rst reported the production of a linear array of regularly spaced silver nanoparticles by using the central channel of wild-type tobacco mosaic virus.89 We have very recently reported the rst example of protein encapsulation into the synthetic glycolipid nanotube from 10.51 The aqueous dispersion of ferritin (5 mg/ mL, pH 6.9) was subjected to the loading experiment of LNTs by capillary action in the similar manner as already mentioned above.47,48 Since iron atoms in the ferrihydrite core can shield the electron beam, TEM image should give a high-contrast image as a black dot for ferritin protein. Figure 14 indicates the TEM image of the glycolipid nanotube of 10 encapsulating ferritins in the hollow cylinder. The presence of ferritin can be conrmed by energy-dispersive X-ray analysis spectrum that shows a Fe Ka peak derived from the ferric core. Considering the inner diameters of $50 nm, at most, four ferritins should align in the normal direction to the long axis of the LNT. Image analysis
Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

allowed us to estimate the lling ratio of ferritin $20%. Recent Progress in Encapsulation The glycolipid nanotube from the glucopyranosylamide lipid 10 also proved to encapsulate magnetite nanocrystals 1030 nm wide into the hollow cylinder.90 In this way, initial lyophilization of the LNTs and subsequent addition of the dried LNTs into the aqueous solution or dispersion containing desired nanocrystals, colloidal particles, and biomacromolecules was found to be very effective to load the nanoscale objects in the LNT hollow cylinder. In addition to the balance between the dimension of guest substances and inner diameters of nanotube hosts, the desired location of appropriate functionalities on the outer and inner surfaces of the nanotube hosts is also critical to separate, deliver, or pattern biomolecules. In particular, the LNTs with positively charged inner surfaces are favorable to encapsulate anionic biopolymers selectively, like DNA, RNA, and certain proteins. We have very recently developed new preparation method to form functional LNTs with a cationic inner surface through the self-assembly

5146

SHIMIZU

Figure 10. Schematic illustration showing a series of experimental procedures to fabricate a gold nanowire via a gold-nanocrystals-encapsulated LNT: (a) lyophilization of the LNT, (b) lling the LNT with a HAuCl4 aqueous solution, (c) photoreduction of Au(III) to Au(0), (d) removal of the outer organic shell, and (e) fusion of gold nanocrystals depending on ring temperature. (Bottom) TEM image of a glycolipid nanotube from 10, lled with gold nanocrystals. [Color gure can be viewed in the online issue, which is available at www.interscience.wiley.com.]

of a novel unsymmetrical bolaamphiphile 15.19 The designed amphiphile possess two hydrophilic groups, glucose and amino moiety at each end, which differ in size and functionalities. Crystalline powders of 15 as self-assembly seeds, which were obtained by reprecipitation from methanol, gave nanotapes on self-assembly. On the other hand, thin lms as seeds, which were prepared by dimethylformamide solution,

self-assembled into mainly nanotube structures with hollow cylinders of 7080 nm. The difference in the initial molecular packing of 15 was found to direct the resultant molecular packing after self-assembly, depending on symmetrical or unsymmetrical MLMs for the crystalline powders or thin lms, respectively, (Fig. 15). All the analytical results obtained supported that the inner most surfaces are exactly covered with
Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

SELF-ASSEMBLED LIPID NANOTUBE HOSTS

5147

Figure 11. TEM images showing different types of connement of gold nanocrystals in the hollow cylinder of the glycolipid nanotube from 10. [Color gure can be viewed in the online issue, which is available at www.interscience.wiley.com.]

amino headgroups charged positively at neutral pH, and the outermost surfaces with 1-glucosamide headgroups. When we attempted the encapsulation of anionic sulfate latex beads 20 nm wide or ferritin proteins in aqueous solutions without depending on the capillary action,47,48 we realized that merely mixing of the aqueous solutions containing the LNT host of 15 and guest substances induced effective encapsulation into the hollow cylinder in solution (Fig. 16). It is already reported that electrostatic interac-

tion plays a key role in coating the outer surfaces or edges of nanotubes selectively with oppositely charged nanoparticles, polymers, and pro-

Figure 12. TGA thermogram of the glycolipid nanotube form 10 encapsulating gold nanocrystals inside the hollow cylinder. Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

Figure 13. FE-SEM images for the assembly of gold nanocrystals after ring the gold-nanocrystals-encapsualted LNT of 10 for 40 min at (a) 450 8C, (b) 550 8C, and (c) 750 8C.

5148

SHIMIZU

teins.9194 The present results suggest that electrostatic interaction is also effective to encapsu-

late biomacromolecules into the hollow cylinder of LNTs.

APPLICATION AS A NANOCHANNEL
When the inner diameters of hollow cylindrical columns or glass capillaries with high-axial ratios are decreasing stepwise from millimeter to nanometer scale via micrometer scale, the number of theoretical plate (NTP) for separation will increase contrary to the decrease in dimensions of the inner diameter. General column chromatography gives ~104 as the NTP. The use of a glass capillary 100 lm wide increases the number to 10 times that of column. If one can

pack brous gels into the capillary, for example, capillary gel electrophoresis reaches to 107 as the NTP,95 the value of which is known to be the present highest level for separation. If we can apply LNTs, the numbers are expected to be 108 or more due to the much smaller sizes of inner diameters. Furthermore, very tiny water volume, which is conned in the hollow cylinder, for example, 10 nm 1000 nm wide (inner diameter length), can accommodate $60 molecules when loading a 1 mM solution. Those high efciencies for separation and the conned effect

Figure 14. (Left) TEM images of ferritin-encapsulated LNTs of 10 and (right) its trace image. Encapsulated ferritin molecules are shown by green circles with brown ferrihydrite cores. [Color gure can be viewed in the online issue, which is available at www.interscience.wiley.com.] Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

SELF-ASSEMBLED LIPID NANOTUBE HOSTS

5149

Figure 15. Plausible molecular packing of 15 for (a) crystalline powder (symmetrical MLMs) and (c) thin lm (unsymmetrical MLMs) as solid seeds, and those for (b) nanotapes and (d) nanotubes as resultant self-assemblies. [Color gure can be viewed in the online issue, which is available at www.interscience.wiley.com.]

for certain guest substances seem to be very potential when applying LNTs as synthetic nanochannels.3 The social and industrial needs to solve the miniaturization of microchannel devices of 50100 lm in diameters have inspired many investigations in supramolecular chemistry. For example, Orwars research group has recently developed a new electroinjection technique that enabled 2D networks of uidstate phospholipids vesicles interconnected by LNTs.7,96 They can also manipulate the connectivity, the container size, the nanotube length, and the angle between the nanotube extensions. These achievements by using LNT architectures contribute to exploit the next generation of biochip devices.43,97101 In particular, appropriately designed nanouidic systems prepared in this manner can provide a useful platform to examine single-molecule dynamics,102 enzyme-catalyzed reactions,103 single-le diffusion,104 and single-molecule sequencing and synthesis.105,106 We have also developed microinjection method to extrude single LNTs of 1 and to place them onto a substrate.107 The moderate stiffness (the Youngs modulus E 720 MPa) evaluated for the single LNT of 1 will allow for the free injection of the nanotube in a \dip-pen" fashion.107 Since this manipulation of the single LNT is carried out in a crystalline state, the present method clearly differs from the Orwars electroinjection technique that treats lipid bilayer systems in a uid phase.97 By using $10 discrete linear arrays of LNTs as a mold on a
Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

glass substrate, we have fabricated a nanochannel device made of resist resin, with each channel having 50100 nm inner diameters.3

CONCLUSIONS
Precisely controlled self-assembly of amphiphilic lipid monomers can give well-dened nanotube architectures with controlled dimensions including inner and outer diameters. Novel molecular design as well as better understanding of molecular self-assembly system enabled us to produce tailor-made-type nanotube hosts with desired functionalities and charges on the inner and outer surfaces. Advanced analytical techniques have unveiled the characteristic properties of water conned in the hollow cylinder of selfassembled LNTs. The noncovalent synthesis of nanotube architectures based on amphiphilic molecules and polymers through self-assembly

Figure 16. TEM image of the LNT from unsymmetrical bolaamphiphile 15, which encapsulate sulfate-latex beads inside the hollow cylinder. The lling of the beads was carried out by mixing two aqueous solutions containing the nanotube and the beads.

5150

SHIMIZU

will further develop the miniaturization and integration of nanostructures in the 10100 nm size region.
The author thanks his colleagues Mitsutoshi Masuda, Hiroyuki Minamikawa, Bo Yang, and Naohiro Kameta for their continuous support and skillful technique at NARC, AIST, during the course of this work on supramolecular lipid nanotubes. George John, Jong Hwa Jung, Shoko Kamiya, Nikolay V. Goutev, and Hiroharu Yui (CREST-JST) are acknowledged for collaboration on the synthesis and analysis of self-assembled tubular architectures. Yoshiki Shimizu (AIST) and Kaname Yoshida (AIST) are gratefully acknowledged for carrying out TEM measurements. Ichiro Yamashita (Matsushita Electric Industrial) is also thanked for collaboration on the encapsulation of ferritin into LNTs. The Japan Science and Technology Agency (JST) is acknowledged for nancial support of the CREST project.

REFERENCES AND NOTES


1. Schnur, J. M. Science 1993, 262, 1669. 2. Block, M. A. B.; Kaiser, C.; Khan, A.; Hecht, S. In Functional Molecular Nanostructures, Schluter, D. A., Ed.; Springer: Berlin, 2005. Topics in Current Chemistry, Vol. 245, pp 89150. 3. Shimizu, T.; Masuda, M.; Minamikawa, H. Chem Rev 2005, 105, 1401. 4. Xu, J.; Li, X.; Liu, J.; Wang, X.; Peng, W.; Li, Y. J Polym Sci Part A: Polym Chem 2005, 43, 2892. 5. Wu, C.; Yang, Y.-T.; Ma, C.-C. M.; Kuan, H.-C. J Polym Sci Part A: Polym Chem 2005, 43, 6084. 6. Huang, H.-M.; Liu, I.-C.; Chang, C.-Y.; Tsai, H.-C.; Hsu, C.-H. J Polym Sci Part A: Polym Chem 2005, 42, 5802. 7. Karlsson, A.; Karlsson, R.; Karlsson, M.; Cans, A.-S.; Stroemberg, A.; Ryttsen, F.; Orwar, O. Nature 2001, 409, 150. 8. Bommel, K. J. C. v.; Friggeri, A.; Shinkai, S. Angew Chem Int Ed 2003, 42, 980. 9. Jung, J. H.; Shinkai, S.; Shimizu, T. Chem Rec 2003, 3, 212. 10. Yager, P.; Schoen, P. E. Mol Cryst Liq Cryst 1984, 106, 371. 11. Yamada, K.; Ihara, H.; Ide, T.; Fukumoto, T.; Hirayama, C. Chem Lett 1984, 1713. 12. Nakashima, N.; Asakuma, S.; Kunitake, T. J Am Chem Soc 1985, 107, 509. 13. Iijima, S. Nature 1991, 354, 56. 14. Iijima, S.; Ichihashi, T. Nature 1993, 363, 603. 15. Spector, M. S.; Easwaran, K. R. K.; Jyothi, G.; Selinger, J. V.; Singh, A.; Schnur, J. M. Proc Natl Acad Sci USA 1996, 93, 12943.

16. Spector, M. S.; Price, R. R.; Schnur, J. M. Adv Mater 1999, 11, 337. 17. Spector, M. S.; Selinger, J. V.; Schnur, J. M. In Topics in Stereochemistry, Vol. 24; Green, M. M.; Nolte, R. J. M.; Meijer, E. W., Eds.; Wiley Interscience: Hoboken, NJ, 2003; pp 281372. 18. Masuda, M.; Shimizu, T. Langmuir 2004, 20, 5969. 19. Kameta, N.; Masuda, M.; Minamikawa, H.; Goutev, N. V.; Rim, J. A.; Jung, J. H.; Shimizu, T. Adv Mater 2005, 17, 2732. 20. Yu, K.; Eisenberg, A. Macromolecules 1998, 31, 3509. 21. Shen, H.; Eisenberg, A. Macromolecules 2000, 33, 2561. 22. Jenekhe, S. A.; Chen, X. L. Science 1999, 283, 372. 23. Stewart, S.; Liu, G. Angew Chem Int Ed 2000, 39, 340. 24. Li, Z.; Liu, G. Langmuir 2003, 19, 10480. 25. Huang, H.; Remsen, E. E.; Kowalewski, T.; Wooley, K. L. J Am Chem Soc 1999, 121, 3805. 26. Kim, Y.; Mayer, M. F.; Zimmerman, S. C. Angew Chem Int Ed 2003, 42, 1121. 27. Martin, C. R. Adv Mater 1991, 3, 457. 28. Martin, C. R. Science 1961, 1994, 266. 29. Steinhart, M.; Wendorff, J. H.; Greiner, A.; Wehrspohn, R. B.; Nielsch, K.; Schilling, J.; Choi, J.; Goesele, U. Science 1997, 2002, 296. 30. Steinhart, M.; Wehrspohn, R. B.; Goesele, U.; Wendorff, J. H. Angew Chem Int Ed 2004, 43, 1334. 31. MacGilivray, L. R.; Atwood, J. L. In Advances in Supramolecular Chemistry, Vol. 6; Gokel, G. W., Ed.; JAI Press: Stamford, CT, 2000; pp 157 183. 32. Biradha, K.; Fujita, M. In Advances in Supramolecular Chemistry, Vol. 6; Gokel, G. W., Ed.; JAI Press: Stamford, CT, 2000; pp 139. 33. Bhattacharyya, K. Acc Chem Res 2003, 36, 95. 34. Scodinu, A.; Fourkas, J. T. J Phys Chem B 2002, 106, 10292. 35. Boyd, J. E.; Briskman, A.; Colvin, V. L.; Mittleman, D. M. Phys Rev Lett 2001, 87, 1474011. 36. Venables, D. S.; Huang, K.; Schmuttenmaer, C. A. J Phys Chem B 2001, 105, 9132. 37. Pant, D.; Levinger, N. E. Langmuir 2000, 16, 10123. 38. Shirota, H.; Horie, K. J Phys Chem B 1999, 103, 1437. 39. Raviv, U.; Laurat, P.; Klein, J. Nature 2001, 413, 51. 40. Zhang, X.; Zhu, Y.; Granick, S. Science 2002, 295, 663. 41. Berger, C.; Desbat, B.; Kellay, H.; Turlet, J. M.; Blaudez, D. Langmuir 2003, 19, 1. 42. Ricci, M. A.; Bruni, F.; Gallo, P.; Rovere, M.; Soper, A. K. J Phys Condens Matter 2000, 12, A345. Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

SELF-ASSEMBLED LIPID NANOTUBE HOSTS

5151

43. Karlsson, A.; Karlsson, M.; Karlsson, R.; Sott, K.; Lundqvist, A.; Tokarz, M.; Orwar, O. Anal Chem 2003, 75, 2529. 44. Kopp, M. U.; Mello, A. J. d.; Manz, A. Science 1998, 280, 1046. 45. Weigl, B. H.; Yager, P. Science 1999, 283, 346. 46. Kenis, P. J. A.; Ismagilov, R. F.; Whitesides, G. M. Science 1999, 285, 83. 47. Yang, B.; Kamiya, S.; Yoshida, K.; Shimizu, T. Chem Commun 2004, 500. 48. Yang, B.; Kamiya, S.; Shimizu, Y.; Koshizaki, N.; Shimizu, T. Chem Mater 2004, 16, 2826. 49. Banerjee, I. A.; Yu, L.; Shima, M.; Yoshino, T.; Takeyama, H.; Matsunaga, T.; Matsui, H. Adv Mater 2005, 17, 1128. 50. Reches, M.; Gazit, E. Science 2003, 300, 625. 51. Yui, H.; Shimizu, Y.; Kamiya, S.; Masuda, M.; Yamashita, I.; Ito, K.; Shimizu, T. Chem Lett 2005, 34, 232. 52. John, G.; Masuda, M.; Okada, Y.; Yase, K.; Shimizu, T. Adv Mater 2001, 13, 715. 53. Yui, H.; Koyama, K.; Guo, Y.; Sawada, T.; John, G.; Yang, B.; Masuda, M.; Shimizu, T. Langmuir 2005, 21, 721. 54. Upadhyay, A.; Bhatt, T.; Tripathi, H. B.; Pant, D. D. J Photochem Photobiol A 1995, 89, 201. 55. Kosower, E. M. Acc Chem Res 1982, 15, 259. 56. Fuhrhop, J.-H.; Wang, T. Chem Rev 2004, 104, 2901. 57. Fuhrhop, J.-H.; Fritsch, D. Acc Chem Res 1986, 19, 130. 58. Fuhrhop, J. H.; David, H. H.; Mathieu, J.; Liman, U.; Winter, H. J.; Boekema, E. J Am Chem Soc 1986, 108, 1785. 59. Schneider, J.; Messerschmidt, C.; Schulz, A.; Gnade, M.; Schade, B.; Luger, P.; Bombicz, P.; Hubert, V.; Fuhrhop, J. H. Langmuir 2000, 16, 8575. 60. Masuda, M.; Shimizu, T. Chem Commun 2001, 2442. 61. Szafran, M.; Dega-Szafran, Z.; Katrusiak, A.; Buczak, G.; Glowiak, T.; Sitkowski, J.; Stefaniak, L. J Org Chem 1998, 63, 2898. 62. Aigouy, P. T.; Costeseque, P.; Sempere, R.; Senac, T. Acta Cystallogr 1995, B51, 55. 63. Masuda, M.; Shimizu, T. Carbohydr Res, 2005, 340, 2502. 64. Singh, A.; Wong, E. M.; Schnur, J. M. Langmuir 2003, 19, 1888. 65. Thomas, B. N.; Corcoran, R. C.; Cotant, C. L.; Lindemann, C. M.; Kirsch, J. E.; Persichini, P. J. J Am Chem Soc 1998, 120, 12178. 66. Thomas, B. N.; Lindemann, C. M.; Corcoran, R. C.; Cotant, C. L.; Kirshch, J. E.; Persichini, P. J. J Am Chem Soc 2002, 124, 1227. 67. Jung, J. H.; John, G.; Yoshida, K.; Shimizu, T. J Am Chem Soc 2002, 124, 10674. 68. Kamiya, S.; Minamikawa, H.; Jung, J. H.; Yang, B.; Masuda, M.; Shimizu, T. Langmuir 2005, 21, 743. Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

69. Jung, J. H.; Shimizu, T. ChemEur J 2005, 11, 5538. 70. Fuhrhop, J. H.; Helfrich, W. Chem Rev 1993, 93, 1565. 71. Fuhrhop, J.-H.; Koening, J. In Monographs in Supramolecular Chemistry; Stoddart, J. F., Ed.; The Royal Society of Chemistry: Cambridge, 1994. 72. Feiters, M. C.; Nolte, R. J. M. In Advances in Supramolecular Chemistry, Vol. 6; Gokel, G. W., Ed.; JAI Press: Stamford, CT, 2000; pp 41156. 73. Helfrich, W.; Prost, J. Phys Rev A: At Mol Opt Phys 1988, 38, 3065. 74. Nelson, P.; Powers, T. Phys Rev Lett 1992, 69, 3409. 75. Ou-Yang, Z.; Liu, J. Phys Rev A: At Mol Opt Phys 1991, 43, 6826. 76. Selinger, J. V.; Schnur, J. M. Phys Rev Lett 1993, 71, 4091. 77. Proulx-Curry, P. M.; Chasteen, N. D. Coord Chem Rev 1995, 144, 347. 78. Tomalia, D. A. In Supramolecular Polymers; Ciferri, A., Ed.; CRC Press: Boca Raton, 2005. pp 187256. 79. Frechet, J. M. J.; Tomalia, D. A. Dendrimers and Other Dendritic Polymers; Wiley: West Sussex, 2001. 80. Bosman, A. W.; Janssen, H. M.; Meijer, E. W. Chem Rev 1999, 99, 1665. 81. Newkome, G. R.; He, E.; Mooreeld, C. N. Chem Rev 1999, 99, 1689. 82. Yang, C. D.; Takeyama, H.; Tanaka, T.; Matsunaga, T. Enzyme Microb Technol 2001, 29, 13. 83. Matsunaga, T.; Kawasaki, M.; Yu, X.; Tsujimura, N.; Nakamura, N. Anal Chem 1996, 68, 3551. 84. Shimizu, T.; Kogiso, M.; Masuda, M. Nature 1996, 383, 487. 85. Kogiso, M.; Ohnishi, S.; Yase, K.; Masuda, M.; Shimizu, T. Langmuir 1998, 14, 4978. 86. Masuda, H.; Hogi, H.; Nishio, K.; Matsumoto, F. Chem Lett 2004, 33, 812. 87. Yamashita, I. Thin Solid Films 2001, 393, 12. 88. Yoshimura, H.; Scheybani, T.; Baumeister, W.; Nagayama, K. Langmuir 1994, 10, 3290. 89. Dujardin, E.; Peet, C.; Stubbs, G.; Culver, J. N.; Mann, S. Nano Lett 2003, 3, 413. 90. Yui, H.; Kamiya, S.; Shimizu, T, unpublished work. 91. Lvov, Y. M.; Price, R. R.; Selinger, J. V.; Singh, A.; Spector, M. S.; Schnur, J. M. Langmuir 2000, 16, 5932. 92. Lvov, Y. M.; Price, R. R. Colloids Surf B: Biointerfaces 2002, 23, 251. 93. Jung, J. H.; Yoshida, K.; Shimizu, T. Langmuir 2002, 18, 8724. 94. Ji, Q.; Iwaura, R.; Kogiso, M.; Jung, J. H.; Yoshida, K.; Shimizu, T. Chem Mater 2004, 16, 250. 95. Guttman, A.; Cohen, A. S.; Heiger, D. N.; Karger, B. L. Anal Chem 1990, 62, 137.

5152

SHIMIZU

96. Karlsson, M.; Nolkrantz, K.; Davidson, M. J.; Stroemberg, A.; Ryttsen, F.; Akerman, B.; Orwar, O. Anal Chem 2000, 72, 5857. 97. Karlsson, M.; Sott, K.; Cans, A.-S.; Karlsson, A.; Karlsson, R.; Orwar, O. Langmuir 2001, 17, 6754. 98. Karlsson, R.; Karlsson, M.; Karlsson, A.; Cans, A.-S.; Bergenholtz, J.; Akerman, B.; Ewing, A. G.; Voinova, M.; Orwar, O. Langmuir 2002, 18, 4186. 99. Karlsson, M.; Sott, K.; Davidson, M.; Cans, A.-S.; Linderhom, P.; Chiu, D.; Orwar, O. Proc Natl Acad Sci USA 2002, 99, 11573. 100. Sott, K.; Karlsson, M.; Pihl, J.; Hurtig, J.; Lobovkina, T.; Orwar, O. Langmuir 2003, 19, 3904.

101. Karlsson, R.; Karlsson, A.; Orwar, O. J Am Chem Soc 2003, 125, 8442. 102. LeDuc, P.; Haber, C.; Bao, G.; Wirtz, D. Nature 1999, 399, 564. 103. Lu, H. P.; Xun, L.; Xie, X. S. Science 1998, 282, 1877. 104. Wei, Q.-H.; Bechinger, C.; Leiderer, P. Science 2000, 287, 625. 105. Sciaky, N.; Presley, J.; Smith, C.; Zaal, K. J. M.; Cole, N.; Moreira, J. E.; Terasaki, M.; Siggia, E.; Lippincott-Schwartz, J. J Cell Biol 1997, 139, 1137. 106. Ishii, Y.; Yanagida, T. Single Mol 2000, 1, 5. 107. Frusawa, H.; Fukagawa, A.; Ikeda, Y.; Araki, J.; Ito, K.; John, G.; Shimizu, T. Angew Chem Int Ed 2003, 42, 72.

Journal of Polymer Science: Part A: Polymer Chemistry DOI 10.1002/pola

You might also like