+ All Categories
Home > Documents > Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf ·...

Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf ·...

Date post: 19-Jul-2020
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
21
N N N M Copyright © 2007 Society of Porphyrins & Phthalocyanines Advances in supramolecularly assembled nanostructures of fullerenes and porphyrins at surfaces Soichiro Yoshimoto* a and Kingo Itaya* b a National Institute of Advanced Industrial Science and Technology (AIST), Central 6, 1-1-1 Higashi, Tsukuba, Ibaraki 305-8566, Japan b Department of Applied Chemistry, Graduate School of Engineering, Tohoku University, 6-6-07 Aoba, Sendai 980-8579, Japan Dedicated to Professor Osamu Ito on the occasion of his retirement Received 15 January 2007 Accepted 26 February 2007 ABSTRACT: The ʻbottom-upʼ strategy is an attractive and promising approach for the construction of nanoarchitectures. Supramolecular assemblies based on non-covalent interactions have been explored in an attempt to control surface properties. In this minireview, we focus on advances made in the past three years in the field of scanning tunneling microscopy (STM) on supramolecular assembly and the function of porphyrins, phthalocyanines, and fullerenes, non-covalently bound on metal single crystal surfaces. Well-defined adlayers, consisting of porphyrin and phthalocyanine for the design of supramolecular nanoarchitectures, supramolecular traps of C 60 on hydrogen bond networks, a unique approach for controlling molecular orientation by a 1:1 supramolecularly assembled film consisting of C 60 and the related derivatives and metallooctaethylporphyrins, and nanoapplications of fullerenes, either induced by tip manipulation or driven by thermal fluctuations at surfaces, were clearly visualized by STM. Copyright © 2007 Society of Porphyrins & Phthalocyanines. KEYWORDS: supramolecular assembly, scanning tunneling microscopy, fullerenes, porphyrins, phthalocyanines, metal single crystal surface, electrochemistry, ultrahigh vacuum. INTRODUCTION Porphyrins provide an extremely versatile synthetic base for a variety of material applications in many disciplines of chemistry and physics, such as opto- electronics, electrochemistry, catalysis, data storage and solar cells [1-8]. It is well-known that porphyrins and related derivatives are included in metallopro- teins as active centers, ʻhemesʼ. To understand their role and electrochemical properties such as dioxygen storage, a research on porphyrin thin films is one of the attractive fields in electrochemistry [9-12]. Since the late 1970ʼs, thin films of metalloporphyrin and metallophthalocyanine derivatives have been studied intensively due to the interest in electrocata- lytic reactions, such as the reduction of O 2 to develop efficient fuel cells; such reactions occur primarily at graphite electrodes [12-16]. However, thus far not much attention has been paid to the adlayer structure of these porphyrins. Formation and characterization of ordered adlayers of porphyrin and phthalocya- nine molecules at electrolyte-electrode interfaces are therefore important in understanding the mechanism of O 2 reduction. On the other hand, covalently linked porphyrin- fullerene C 60 dyads were used as one of the artificial photosynthetic models in the mid 1990ʼs [17- 19]. Because fullerenes are known as one type of acceptor with strong π-electron accepting ability, they are considered to be suitable building blocks *Correspondence to: Soichiro Yoshimoto, email: so- [email protected], fax: +81 29-861-6177 and Kingo Itaya, email: [email protected], fax: +81 22- 795-5868 Journal of Porphyrins and Phthalocyanines Microreview J. Porphyrins Phthalocyanines 2007; 11: 313-333 Published at http://www.u-bourgogne.fr/jpp/ Published on web 05/30/2007
Transcript
Page 1: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

N

NN

M

Copyright © 2007 Society of Porphyrins & Phthalocyanines

Advances in supramolecularly assembled nanostructures of fullerenes and porphyrins at surfaces

Soichiro Yoshimoto*a and Kingo Itaya*b

a National Institute of Advanced Industrial Science and Technology (AIST), Central 6, 1-1-1 Higashi, Tsukuba, Ibaraki 305-8566, Japan b Department of Applied Chemistry, Graduate School of Engineering, Tohoku University, 6-6-07 Aoba, Sendai 980-8579, Japan

Dedicated to Professor Osamu Ito on the occasion of his retirement

Received 15 January 2007Accepted 26 February 2007

ABSTRACT: The ̒ bottom-up ̓strategy is an attractive and promising approach for the construction of nanoarchitectures. Supramolecular assemblies based on non-covalent interactions have been explored in an attempt to control surface properties. In this minireview, we focus on advances made in the past three years in the field of scanning tunneling microscopy (STM) on supramolecular assembly and the function of porphyrins, phthalocyanines, and fullerenes, non-covalently bound on metal single crystal surfaces. Well-defined adlayers, consisting of porphyrin and phthalocyanine for the design of supramolecular nanoarchitectures, supramolecular traps of C60 on hydrogen bond networks, a unique approach for controlling molecular orientation by a 1:1 supramolecularly assembled film consisting of C60 and the related derivatives and metallooctaethylporphyrins, and nanoapplications of fullerenes, either induced by tip manipulation or driven by thermal fluctuations at surfaces, were clearly visualized by STM. Copyright © 2007 Society of Porphyrins & Phthalocyanines.

KEYWORDS: supramolecular assembly, scanning tunneling microscopy, fullerenes, porphyrins, phthalocyanines, metal single crystal surface, electrochemistry, ultrahigh vacuum.

INTRODUCTIONPorphyrins provide an extremely versatile synthetic

base for a variety of material applications in many disciplines of chemistry and physics, such as opto-electronics, electrochemistry, catalysis, data storage and solar cells [1-8]. It is well-known that porphyrins and related derivatives are included in metallopro-teins as active centers, ʻhemesʼ. To understand their role and electrochemical properties such as dioxygen storage, a research on porphyrin thin films is one of the attractive fields in electrochemistry [9-12]. Since the late 1970ʼs, thin films of metalloporphyrin

and metallophthalocyanine derivatives have been studied intensively due to the interest in electrocata-lytic reactions, such as the reduction of O2 to develop efficient fuel cells; such reactions occur primarily at graphite electrodes [12-16]. However, thus far not much attention has been paid to the adlayer structure of these porphyrins. Formation and characterization of ordered adlayers of porphyrin and phthalocya-nine molecules at electrolyte-electrode interfaces are therefore important in understanding the mechanism of O2 reduction.

On the other hand, covalently linked porphyrin-fullerene C60 dyads were used as one of the artificial photosynthetic models in the mid 1990ʼs [17-19]. Because fullerenes are known as one type of acceptor with strong π-electron accepting ability, they are considered to be suitable building blocks

*Correspondence to: Soichiro Yoshimoto, email: [email protected], fax: +81 29-861-6177 and Kingo Itaya, email: [email protected], fax: +81 22-795-5868

Journal of Porphyrins and Phthalocyanines MicroreviewJ. Porphyrins Phthalocyanines 2007; 11: 313-333 Published at http://www.u-bourgogne.fr/jpp/

Published on web 05/30/2007

Page 2: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

S. YOSHIMOTO AND K. ITAYA314

for the construction of three-dimensional molecular architectures [20]. Subsequently, the non-covalent interaction between C60 and porphyrin cyclic dimer [21, 22] and C60 and palladium-linked bis-porphyrin was reported [23]. Currently, fullerene-porphyrin supramolecular assemblies have been studied exten-sively to generate photocurrents as well as to eluci-date their unique photophysical and photochemical properties [2-8, 24, 25]. Since establishing the concept for supramolecular chemistry [26-28], self-assembly of organic molecules has recently gained conside-rable attention as a key technique for the ʻbottom-up ̓fabrication of nano-scale functional structures [26-33]. For example, self-assembled porphyrin trimers and graphitic nanotubes are attractive materials for surface patterning [34] and molecular electronics [35]. Research on programmed DNA assembly, the so-called ʻDNA-origami ̓ [36] and ʻDNA-nanotile ̓[37-40], has also been an attractive topic. A new concept of biosensing to achieve multiplexed de-tection was proposed by using combinatorial self-assembly of DNA-nanotiles to direct the self-as-sembly of fluorescently labeled molecular probes into combinatorial encoding arrays [40].

In recent years, scanning tunneling microscopy (STM) and atomic force microscopy (AFM) have been widely accepted as powerful tools for under-standing the structure of adsorbed layers of mole-cules on metal surfaces at the atomic scale, both in ultrahigh vacuums (UHV) [41-44] and in aqueous solutions [45-51]. High-resolution STM has made it possible to directly determine packing arrangements and even internal structures of adsorbed organic mole-cules. Furthermore, it has been demonstrated that in situ STM makes it possible to monitor, with atomic or molecular resolution, a wide variety of electrode processes such as the adsorption of water-soluble inorganic and organic species and underpotential de-position of metal ions [45-48]. For the construction of supramolecular assembled nanoarchitectures, the appropriate modifications for concepts of solution-based supramolecular chemistry are necessary to further extend supramolecular assembly technique on substrates. Therefore, molecular level visualization is very important to fully understand, not only the interaction between substrate and adsorbates, but also the influence of electronic properties of adsorbed ligands on non-covalent bonds. Supramolecular as-semblies based on non-covalent interactions such as dipole-dipole, hydrogen bonding and metal-ligand coordination on metal surfaces, have also been ex-plored in an attempt to control surface properties [43, 44].

In this minireview, we focus on the molecular assemblies of functionalized molecules such as porphyrins, phthalocyanines and fullerenes, non-covalently bound on metal single crystal surfaces,

because those molecules are important building blocks for the construction of nanostructures. This article primarily focuses on our own experimental results, obtained on Au single crystal electrodes in aqueous electrolyte solutions under electrochemical conditions, in addition to several published results from other groups in the last three years.

Porphyrin and phthalocyanine adlayers

Single component adlayers. Adlayer structures of porphyrin and phthalocyanine derivates have been previously studied, mostly in UHV using STM on various metal surfaces [52-60]. Lippel et al. reported the first STM images of a copper(II) phthalocyanine (CuPc) adlayer on Cu(100) [52]. Gimzewski and coworkers investigated copper(II) 5,10,15,20-tetra-kis(3,5-di-t-butylphenyl)porphyrin (CuTBPP) on Cu(100), Au(110), and Ag(110) in UHV [53, 54] and found that the packing arrangement of CuTBPP depends on the metal substrate used. Hipps and co-workers reported various MPcʼs (M; Cu [55, 56], Co [55, 56], Ni [57], Fe [57], VO [58]) and metallotetra-phenylporphyrins (MTPP) [59, 60] on reconstructed Au(111) and found that the brightness of the center spot of Pc or TPP is dependent on the active center metal. The difference in contrast between the metal ions in STM images was explained in terms of the occupation of the dz2 orbital. As listed above, STM imaging of hydrophobic MPc or MTPP molecules was successful in UHV.

STM has also been utilized in understanding the structure of adlayers of water-soluble porphyrin mo-lecules in aqueous solutions [46, 61-70]. Highly orde-red arrays of free-base 5,10,15,20-tetrakis(N-methyl-pyridinium-4-yl)porphyrin (H2TMPyP) molecules were formed on the iodine-modified (I-) Au(111) [61-63], I-Ag(111) [64], I-Pt(100) [65], and S-modified Au(111) [66] electrodes. However, the ordering of a free-base porphyrin array directly attached to a bare Au(111) surface at an electrochemical interface was subsequently reported by He et al. [67, 68]. Potential manipulation plays a significant role in controlling the surface mobility of tetrakis(4-pyridyl)porphyrin (H2TPyP) molecules [67]. Tao et al. also investi-gated the adlayers of three water-soluble molecules, iron(III) protoporphyrin, zinc(II) protoporphyrin, and protoporphyrin(IX), on the graphite basal plane in aqueous solutions with both STM [69, 70] and AFM [69]. Similar adlayer structures were formed using these three molecules, although the internal structures obtained by in situ STM were significantly different [69].

We subsequently investigated several water-inso-luble porphyrin adlayers such as cobalt(II) porphine (CoP) [71], cobalt(II) octaetylporphyrin (CoOEP) [71], cobalt(II) tetraphenylporphyrin (CoTPP) [72],

Page 3: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

SUPRAMOLECULARLY ASSEMBLED NANOSTRUCTURES 315

cobalt(II) coordinated ʻpicket-fence ̓ porphyrin (CoTpivPP) [73] (these molecules are abbreviated as CoPor), and cobalt(II) phthalocyanine (CoPc) [74] (see Chart 1). We were successful in forming highly ordered molecular arrays of those molecules sponta-neously on Au(111) surfaces by immersing Au(111) in benzene solutions containing those molecules. The enhancement of reductive current for O2 reduction at those modified Au(111) electrodes compared to the bare Au(111) electrodes clearly shows that CoPor and CoPc adlayer catalyzes the reduction of O2. By using rotating CoPor- or CoPc-modified Au(111) disk electrodes, it was indicated that the 2-electron reduction process of O2 to H2O2 proceeded on

the CoP-, CoTPP-, CoOEP-, and CoPc-modified Au(111) surfaces [50, 71-74]. High-resolution STM images for each adlayer on Au(111) are shown in Fig. 1. CoP formed highly ordered arrays following a slightly negative potential manipulation (see Fig. 1a). For the other adlayers, such as CoOEP, CoTPP and CoPc molecules, stable and highly ordered arrays were formed on the Au(111) surface. The ad-layers were not dependent upon electrode potential at or near the H2 evolution potential in HClO4. The adlayer structures of CoOEP and CoTPP formed in benzene solutions were identical to those obtained in UHV [60, 75]. The CoOEP and CoTPP adlayers formed an incommensurate structure with respect to

Chart 1. Chemical formulae of CoP, CoOEP, CoTPP, CoTpivPP, CoTCPP, CoCTPP, CoPc and CRCoPc

Fig. 1. Typical high-resolution STM images (9 × 9 nm2) of (a) CoP, (b) CoOEP, (c) CoTPP, (d) CoTpivPP, (e) CoTCPP, (f) CoCTPP, (g) CoPc and (h) CRCoPc adlayer on Au(111) in 0.1 M HClO4, respectively. Reprinted with permissions from references 71-74, copyrights © 2003, 2004 and 2007 American Chemical Society and from references 76, 77, copyrights © 2003 and 2006 The Royal Society of Chemistry, respectively

Page 4: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

S. YOSHIMOTO AND K. ITAYA316

the reconstructed Au(111) surface. For the CoTpivPP, individual CoTpivPP molecules were recognized to be square-shaped with four bright spots at the corners in the characteristic nanobelt array, whereas the ad-lattice was almost the same as that of CoTPP. Care-ful inspection revealed the presence of three small protrusions separated from each other in each bright spot, suggesting that they are tert-butyl groups in each CoTpivPP molecule, as marked by white circles in Fig. 1d. The brightest one was assigned to the O2-coori-nated CoTpivPP molecule. The state of O2 trapped in the cavity of CoTpivPP was distinctly observed in STM images as a bright spot in the nanobelt array formed on reconstructed Au(100)-(hex) surface, but not on the Au(111) surface, indicating that the forma-tion of nanobelt arrays consisting of O2-adducted CoTpivPP molecules was dependent upon the crystal-lographic orientation of Au [73]. Furthermore, we in-vestigated the organization and hydrogen bonding net-work formation of carboxyphenyl porphyrin deriva-tives at electrochemical interfaces [76]. Our results demonstrate that the ordered arrays of monocarboxy- or tetracarboxy-substituted porphyrin derivatives for-med on a Au(111) surface in acidic solution depend, to a remarkable extent, on the number of peripheral substituents (i.e. dimer and tetramer structures were formed for monocarboxy- or tetracarboxy-substituted porphyrin derivatives, respectively, by the variation of the applied potentials, see Fig. 1e,f). Electrochemi-cal control of the self-assembly of porphyrin com-plexes is an effective method for constructing novel surface architectures. In the case of CoPc, the packing arrangement of CoPc on the reconstructed Au(111) surface was also consistent with that obtained in UHV by Lu et al. [55, 56]. However, we found other adlayer structures following preparation from the solution phase. CoPc molecules formed three packing arrangements on Au(111) surface, i.e. one rectangular arrangement on reconstructed Au(111) and two hexagonal arrangements on Au(111)-(1 × 1), depending on the surface coverage [74]. Figure 1g shows a typical STM image for the CoPc adlayer with (3√3 × 3√3)R30˚ symmetry on Au(111)-(1 × 1) surface. The method of immersion into benzene solu-tion was further extended to a larger molecule such as 15-crown-5-ether-substituted cobalt(II) phthalo-cyanine (CoCRPc) [77]. Each molecule was seen as propeller-shaped with the brightest spot at the center, with four additional bright rings at the corners of each CoCRPc molecule (see Fig. 1h). The adlattice of the CoCRPc array on Au(111) was determined to be p(8 × 4√3R-30˚). In the absence of Ca2+, four additional spots were observed at the corners, whereas in the presence of Ca2+, only two additional bright spots were observed at diagonal positions with respect to the Pc ligand, as a result of encapsulation of two Ca2+ ions by crown ether rings [77]. Furthermore, it was

demonstrated by using Au(100)-(1 × 1) lattice that the relationship between crown moieties of CRPc and the underlying Au lattice is important in the trapping of Ca2+ ions in crown rings [78].

In contrast, adlayers containing other central me-tal ions, for example, CuTPP, NiTPP, ZnTPP, CuOEP, NiOEP, ZnOEP, CuPc, and ZnPc, were observed as a dark spot in the center both in UHV [55-60] and in solution [72, 74, 79-82]. Molecular resolution STM revealed a propeller shaped ad-molecule or circle shape with eight additional spots at the corners with its center imaged as a protrusion for CoII and a de-pression for other metal ions. The CoPc or CoTPP molecules were easily identified by the strong tun-neling current resulting from orbital mediated tun-neling through the half-filled dz2 orbital of the CoII ion (d7): a bright spot appeared at the center of each molecule, whereas NiII (d8), CuII (d9) ZnII (d10) ions have a fully filled dz2 orbital [55-60, 81]. The remarkable contrast in tunneling current afforded by the difference in electronic configuration of transition metal ions allows one to clearly distinguish between species for chemical identification at the molecular level. In the case of ZnII ion, we found the possibility of epitaxial growth of ZnPc layer from the solution phase. ZnPc molecules can be epitaxially assembled because of the attractive interaction between the elec-tron donating pair in the Pc scaffold and the zinc ion. The formation of epitaxial layers of ZnPc molecules from benzene solutions was controllable by altering the immersion time [81].

Direct synthesis of metalloporphyrin at Ag(111) was reported by Gottfried et al. [83, 84]. This group prepared free base H2TPP adlayer on Ag(111) in UHV, then Co was further deposited at the adlayer. XPS results indicated that Co ion can be coordinated to a porphyrin ring [83]. This method was effective for Zn-metalation to H2TPP adlayer by Zn deposition [84]. Very recently, direct observation of iron-meta-lation into a free-base porphyrin array were inde-pendently reported by two groups [85, 86]. Auwärter and coworkers succeeded in directly observing Fe-metalation at H2TPyP array on Ag(111) in UHV [85]. Buchner et al. also showed iron-metalation with H2TPP array formed on Ag(111) by using UHV-STM [86]. In both cases, each H2TPyP or H2TPP molecule appeared as brighter spots in the high-resolution STM image after the deposition of iron on those adlayers, suggesting that the metalation of the Fe atom occur-red at the H2TPyP and H2TPP adlayers on Ag(111). The precise control of function (such as dioxygen reduction activity) is determined by the method of direct coordination of various metal ions.

Structure-controlled binary arrays. Porphyrins are also well-known as suitable building components to construct two- or three-dimensional self-organized

Page 5: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

SUPRAMOLECULARLY ASSEMBLED NANOSTRUCTURES 317

multiarrays [87-94]. Self-assembly of complementary subunits via non-covalent interactions such as hy-drogen bonds and metal-ligand coordination is an attractive approach to the controlled oligomeri-zation of monomers to form large supramolecular architectures [91-94]. The binary system consisting of porphyrin and phthalocyanine will aid in under-standing the supramolecular design of nanoarchi-tectures. Hipps and coworkers found the formation of well-ordered regions with an entirely new structure with 1:1 composition of cobalt(II) hexadecafluoro phthalocyanine (F16CoPc) and NiTPP on Au(111) by vapor-phase deposition in UHV [95]. F16CoPc and NiTPP molecules were distinguished from each other by the difference in brightness of central metal ions between F16CoPc and NiTPP molecules. The remarkable contrast in tunneling current afforded by the difference in electronic configuration of transi-tion metal ions allows one to clearly discriminate between species for chemical identification at the molecular level. When CoPc and NiTPP molecules were deposited on Au(111), the binary adlayer consisting of CoPc and NiTPP revealed a densely packed, apparently well-defined structure, which is compositionally disordered [96]. Hipps et al. concluded that the formation of 1:1 composition of F16CoPc and NiTPP is explained by the ability of the two types of molecules to interlace due to the up-to-4 kJ.mol-1 attractive energy for each close-approach fluorine-hydrogen intermolecular interaction and the reduced repulsive interaction between F16CoPc mo-lecules because of the increase of spacing forced by NiTPP units. They also investigated CoPc and CoTPP mixed system on Au(111). The electronic structure (state) between CoPc and CoTPP was differentially identified in the alternate binary CoPc and CoTPP array by STM and the orbital-mediated tunneling spectra measurements [97].

Another system, a bimolecular array consisting of CoPc and CuTPP, was examined both on the

Au(111) and Au(100) surfaces by immersing the sur-faces in benzene solutions [98, 99]. We found that an alternate mixed layer consisting of CoPc and CuTPP was formed on reconstructed Au(100)-(hex) but not on reconstructed Au(111), suggesting that the supramolecular assembly consisting of two chemical components also depends upon the crystallographic orientation of Au [98]. In this case, the terrace of Au(111) was completely covered with two different phases: a disordered region and a highly ordered re-gion consisting of CuTPP, whereas the stripes com-posed of alternate bright and dark lines were obser-ved on Au(100)-(hex) (see Fig. 2a). The difference in brightness at the centers of CoPc and CuTPP molecules is clearly explained by the difference in the mode of occupation of d orbitals, as described by Hipps and coworkers [55, 56, 96]. Subsequently, we found that a similar two-component supramolecular adlayer consisting of CoPc and CuTPP was formed on Au(111) following the different modification con-dition [99]. It is likely that the adsorbate-substrate interaction is much weaker on Au(111) than on Au(100)-(hex). The formation of the stripe structure was independent of the crystallographic orientation, while the corrugation periodicity of reconstruction for the underlying substrate is quite different from Au(111) and Au(100). One possibility is that the difference in π-electron donating ability between CoPc and CuTPP might play a significant role for the alternate stripe formation. We believe that the molecular assembly in this system is somewhat weak and dominated by van der Waals interactions. A bimolecular system consisting of CuOEP and CoPc on Au(111) was also examined by our group at the electrochemical interface [100]. In the case of the bimolecular array consisting of CuOEP and CoPc, it has a p(9 × 3√7R-19.1˚) and its mirror structure, p(9 × 3√7R-40.9˚), were alternately formed on the Au(111) terrace. Furthermore, we attempted to con-trol the domain size and composition by changing

the potential scan at a slow rate to a sligh-tly cathodic potential. In this case, one-dimensional (1D) molecular chains of CuOEP were clearly observed as dark gaps between bright rows consisting of two or three CoPc molecular rows, as shown in Fig. 2b, with the result that the structure chan-ged drastically upon this potential manipu-lation [100]. In this system, the surface mobility and the molecular reorganization of CuOEP and CoPc were accelerated by variation of electrode potential. The sur-face charge density at the electrochemical interface would contribute not only to the interaction between molecule and substrate but also to the interaction between mole-cules. Such a precise and unique control

Fig. 2. Typical STM images of the binary adlayers consisting of (a) CoPc and CuTPP on the reconstructed Au(100)-(hex) surface and (b) CoPc and CuOEP on Au(111) induced by potential variation in 0.1 M HClO4. Reprinted with permission from reference 99, copyright © 2006 American Chemical Society

Page 6: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

S. YOSHIMOTO AND K. ITAYA318

of electrochemical interfaces is of great interest for exploring further applications of the porphyrin and phthalocyanine molecular assembly.

Double- and triple-deckers. In recent years, ad-layers of several sandwich compounds such as the bis-phthalocyanine double-decker complexes of Y, Ce, Pr La, and Er, have been investigated at the liquid-solid interface (mainly on HOPG in 1-phenyloctance) [101-107]. Possible future applications for these rare earth complexes include their use as organic field-effect transistors, liquid crystals, chemical sensors [108] and especially, molecular memory of surfaces, as has been proposed by Lindsey and coworkers [109]. In the case of double-decker heterodimers, several groups have reported a unique property of molecular rotation [110-113]. This rotation can be controlled by coordination to specific functional groups within the molecule [110-112] or by the redox properties of the rare earth central metal [113]. Thus, it is very important to understand the formation of double- or triple-decker adlayers by rotation control

and electrochemical properties. We very recently reported the adlayer of 15-crown-5-ether-substituted Pc and tetraphenyl porphyrin (TPP) triple-decker complex coordinated by EuIII, TPP/CRPc/CRPc on a Au(111) surface to create a three-dimensional functional molecular architecture [114]. A unique molecular assembly and new surface properties of TPP/CRPc/CRPc was found as a result of the adsorption orientation and electrochemical potential variation on the Au(111) surface. As can be seen in Fig. 3a,b, a characteristic well-ordered domain of 2:1 sandwich complex consisting of phthalocyanines and tetraphenylporphyrin was found on a Au(111) surface in HClO4 by in situ STM, indicating that a highly ordered array was formed by alternately arranging two different molecular orientations, as illustrated in Fig. 3c. This finding is applicable to “bottom-up” nano-fabrication of the surface by controlling the molecular orientation and the rotation of the ligands.

Fig. 3. (a) Typical STM image (30 × 30 nm2) and (b) a height-shaded plot of TPP/CRPc/CRPc molecular array formed on Au(111) obtained at 0.80 V vs. RHE in 0.05 M HClO4. (c) Proposed structural model of an alternately arranged superstructure of in the adlayer of TPP/CRPc/CRPc. Reprinted with permission from reference 114, copyright © 2007 Wiley-VCH

(c)

Page 7: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

SUPRAMOLECULARLY ASSEMBLED NANOSTRUCTURES 319

Supramolecular assembly of fullerenes at sur-faces

Host-guest interface using hydrogen bonding architectures. Several interesting nanoarrays were also found in UHV environment. Hydrogen bond-based network structures provide a potential path-way to the design of host-guest interfaces, because the cavity size can be controlled through careful selection of the component molecules. For example, the formation of a self-assembled bimolecular network through hydrogen bonds by co-adsorption of perylene tetra-carboxylic di-imide (PTCDI) and 1,3,5-triazine-2,4,6-triamine (melamine) on Ag/Si(111)-(√3 × √3)R30˚ in UHV was reported by Theobald et al. [115]. The substrate, Ag deposited on Si(111), is one of the suitable surfaces for adsorption of organic molecules such fullerenes, phthalocyanine, and naphthalene tertracarboxylic di-imide (NTCDI) in UHV environment, as those molecules diffuse freely on the deposited surface and form islands in which the order is predominantly governed by inter-molecular interactions. The framework provides a regular array of identical nanoscale traps in which

further deposited molecules nucleate cluster growth. Such network arrays could recognize selectively fullerenes C60 or C84 as guest molecules [115, 116] and it was demonstrated that the seven C60 molecules were included in a 2D open honeycomb network whereas three C84 molecules were trapped in the cavity. According to a recent study on adlayers composed of those components, a similar host-guest function was also found on Au(111) [117]. The bimolecular network consisting of melamine and PTCDI on Au(111) was formed by elevating the temperature (see Fig. 4a,b). Supramolecular traps of C60 molecules occurred at the template of the bimolecular network consisting of melamine and PTCDI. The cluster size of C60 depends upon the coverage of C60. When the coverage is lower than 0.1 monolayers, an open hexamer is predominantly formed in the bimolecular network array, as shown in Fig. 4c. At higher co-verage, the hexamers transform to heptamers and C60 molecules attach over the top of the bimolecular network through further C60 deposition (see Fig. 4d), as illustrated in Fig. 4e.

Reutt-Robey and coworkers reported that a uni-que supramolecular assembly of C60 on the adlayer of acridine-9-carboxylic acid (ACA) formed on Ag(111) in UHV [118, 119]. Because ACA molecules can form several adlayers depending on the coverage on Ag(111) [119], the adlayer structure of C60 in the supramolecular assembled layer is also dependent on the ACA adlayers. For example, when 0.65 ML of C60 is added to the ACA adlayer of the initial coverage of 0.3 ML, adlayers with chiral structures and linear arrangements of C60 were predominantly formed on Ag(111) as shown in Fig. 5a. Close-up views of C60 arrays on chiral and linear structures are shown in the high-resolution STM images of Fig. 5b,c. Each C60 molecule positions on the central portion of each ACA trimer in the chiral structure consisting of alternately arranged ACA trimers as a “pinwheel” shape (with the (2√21 × 2√21R ± 10.9˚) symmetries), whereas C60 molecules are located on the linearly-arranged ACA adlayer. The subtle control of the coverage plays a significant role in the formation of supramolecular pattern.

It is a very attractive and effective method to use hydrogen bonding for two-dimensional networks or arrays of porphyrins at surfaces. For example, Lei et al. reported that high-ly ordered arrays of 5,10,15,20-tetrakis(4-carboxyphenyl)porphyrin (H2TCPP) and copper(II) 2,3,9,10,16,17,23,24-octakis(car-boxyl)phthalocyanine (CuPc8C) were for-

Fig. 4. STM images and proposed structures of (a and b) PTCDI-melamine bimolecular network formed on Au(111) and of (c-e) supramolecular traps of C60 on the bimolecular network observed in UHV. Reprinted with permission from reference 117, copyright © 2006 American Chemical Society

Page 8: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

S. YOSHIMOTO AND K. ITAYA320

med on HOPG by the coadsorption of stearic acid and 1-iodooctadecane, respectively, under ambient conditions [120]. Because 2D hydrogen bonds of porphyrin and Pc were stabilized by the presence of those alkane derivatives, the alkane derivatives contribute the minimization of surface free energy in a 2D system. Furthermore, as reported by Otsuki et al., 5,10-bis(4-carboxyphenyl)-15,20-bis(4-octadecyl-oxyphenyl)porphyrin was found to form dimer rows at the dichlorobenzene-HOPG interface, suggesting that the adlayer structure is determined by directional hydrogen-bonding interactions between nearest neighbor molecules [121].

In contrast, on Au(111), the supramolecular nano-structures composed of carboxyphenyl-substituted porphyrins can be precisely controlled by the con-formation selective assembly [122]. Supramolecular assembly using porphyrin derivatives has also

been investigated to potentially fabricate precisely-controlled molecular wires. Yokoyama and coworkers showed that selective aggregation of porphyrin derivatives are rationally controlled by tuning the dipole-dipole interactions between molecules [123]. They found that CN-substituted (3,5-di-t-butyl-phenyl)porphyrin derivative (TBPP) molecules could produce supramolecular aggregates such as trimers, tetramers, and one-dimensionally extended wire-like structures on Au(111) at low temperatures (63 K) in UHV. Dipole-dipole interactions between CN moieties is a key factor for molecular assembly of TBPP. A similar selective assembly was also formed by carboxyphenyl porphyrins [122]. A remarkable difference between CN- and COOH-moieties in TBPP derivatives is the variation of intermolecular inte-raction between porphyrin derivatives. For example, one-carboxyphenyl-substituted TBPP molecules

form a dimer structure through hydrogen bonding between two molecules (see Fig. 6a), whereas one CN-substituted TBPP mo-lecules arrange in clusters of trimers. The supramolecular selective assembly of 5,15-bis(4-carboxyphenyl)-10,20-bis(3,5-di-t-butylphenyl)porphyrin (trans-BCaTBPP) with a long, straight wire was found on the Au(111) surface by sequential hydrogen bonding between carboxyphenyl groups (see Fig. 6b), while the isolated supramole-cular wires were formed by cyanophenyl-substituted TBPP molecules [123]. In the case of cis-BCaTBPP, the supramolecular assemblies depended upon the coverage. At low coverage, cis-BCaTBPP forms tetramers the same as cis-CN-substituted TBPP molecules, as shown in Fig. 6c. Increasing the coverage, the condensed structure of cis-BCaTBPP composed of a zigzag arrangement appeared on the Au surface. The supramolecular nanostructures on the surfaces can be precisely controlled by the conformation selective assembly. In addition, it was very recently reported that linear C60 arrays were formed on the supramolecular assembled trans-BCaTBPP wires on Au(111) [124]. As reported by the same authors, trans-BCaTBPP forms linearly arrays on Au(111), as shown in Fig. 6d. However, linear arrays of opened nanopore structures consisting of trans-BCaTBPP molecules were formed, i.e. a lateral shift of the supramolecular wires which is accompanied by a partial change in the conformation of trans-BCaTBPP was caused by deposition of C60 molecules. High-resolution STM images revealed that each C60 molecule was located on the open

Fig. 5. (a) Large-scale STM image of C60 arrays on the ACA (coverage 0.75ML) adlayer on Ag(111). High-resolution STM images for (b and c) and (d and e) proposed models of C60 arrays on chiral domain and linear domain, respectively. Reprinted with permission from reference 119, copyright © 2006 American Chemical Society

Page 9: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

SUPRAMOLECULARLY ASSEMBLED NANOSTRUCTURES 321

nanopores, not on the center of each trans-BCaTBPP molecule (see Fig. 6e,f), suggesting that the opening of the nanopores results from the subtle balancing of substrate-molecule and molecule-molecule interactions.

Lin and coworkers reported the construction and control of various surface structures such as chirality, nano cavity, and network array by metal-organic coordination in UHV [44, 125-132]. They reported that terephthtalic acid (TPA), 1,2,4-tricarboxylic benzoic acid (TMLA), 4,1ʼ,4ʼ,1ʼ̓ -terphenyl-1,4ʼ̓ -dicarboxylic acid (TDA) molecules easily undergo complexation with coadsorbed Fe [125, 126]. For example, a two-dimensional nanoporous network was formed on Cu(100) by the metal-organic coor-dination either between TPA and Fe or between TDA and Fe [125]. The tuning of 2D metal-organic networks using Fe and TPA or TDA is shown in STM images in Fig. 7a-c. An adlayer structure consisting of Fe and TPA was dependent upon the concentration of Fe, i.e. ladder-type structure with elongated cavities was dominantly formed on Cu(100) with low concentration of Fe (Fig. 7a), whereas a trellis with a square unit cell and a cross-shaped cavity was found on complete 2D Fe-carboxylate reticulation (Fig. 7b). In the case of TDA, a higher analogue to networks obtained with TPA can provide the construc-tion of 2D metal-organic coordination (Fig. 7c). Such network arrays composed of metal atoms and organic molecules could selectively recognize C60 molecules as a guest molecule. As seen in the STM image shown in Fig. 7d, single C60 (yellow spheres) could be linearly arranged on a ladder-type network with

preferential occupation of available larger cavities (C60 clusters are on areas with bare Cu substrate). The network cavities consisting of Fe and TPA accommodated single C60 molecules, whereas either dimers or trimers of C60 molecules were trapped in single cavities composed of Fe and TDA. This method has been applied to Co and TPA network formation on Au(111) [130, 131]. In addition, the symmetry of the evolving coordination networks is independent of the difference in metal substrate and in crystallo-graphic orientation, which indicates that the metal-ligand coordination predominates over the substrate. The Co-coordination, which is rarely encountered in 3D metal-organic frameworks, is promoted on Cu(100) and Ag(111) by the strict confinement of a 2D environment [132]. Thus, selection of coordination metal, iron or cobalt, makes it possible to precisely control this kind of nanostructure.

Fullerene and porphyrin supramolecular assem-bly. It is known that porphyrins and fullerenes spon-taneously attach to each other [25]. For example, MOEP and C60 form a supramolecular assembly through the π-π interaction in cocrystallites with a ratio of 2 : 1 [133]. To control supramolecular assem-blies between porphyrins and fullerenes during three-dimensional construction, layer-by-layer growth on metal surfaces must be used as a first step. Such a supramolecular assembly produced through π-π or donor-acceptor interaction would be useful for the design and organization of functional organic molecules on electrode surfaces. For example, individual CuPc molecules were observed on a highly

ordered C60 array formed on Au(111) in UHV [134]. Subsequently, we first reported a supramolecular assembled array of C60 molecules on ZnOEP adlayer on Au(111) through a wet process, i.e. successive immersion into benzene solution containing ZnOEP and C60 molecules and the adlayers were observed under the electrochemical environment [80]. As reported in several papers, the thin epitaxial film of C60 on Au(111) was found to take two different close-packed structures, (2√3 × 2√3)R30˚ and the so-called “in-phase” (38 × 38), not only in UHV [135-137] but also in solution [138, 139]. The intermolecular distance between C60 molecules was 1.0 nm in the adlayer of C60-directly attached Au(111), whereas the intermolecular distances between the nearest neighbor C60 molecules were found to be either 1.65 ± 0.07 or 1.40 ± 0.05 nm for directions of two molecular rows on the ZnOEP adlayer [80]. The adlayer structure of ZnOEP on Au(111) was identical to that of CoOEP on

Fig. 6. Large-scale (25 × 40 nm2) STM images and structural models of (a) CaTBPP, (b) trans-BCaTBPP, and (c) cis-BCaTBPP at low-coverage formed on Au(111) in UHV at 63 K. Reprinted with permission from reference 122, copyright © 2004 American Institute of Physics. (d) Large-scale (25 × 35 nm2) and (e) high-resolution STM images and (f) proposed model of C60 array formed on the trans-BCaTBPP adlayer on Au(111). Reprinted with permission from reference 124, copyright © 2007 Wiley-VCH

Page 10: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

S. YOSHIMOTO AND K. ITAYA322

Au(111). The fact that the intermolecular distances between C60 molecules are nearly equal to the dis-tance between ZnOEP molecules indicates that each C60 is located on the center above each ZnOEP mole-cule [50]. The identical structures of C60 and C70 were also observed on a NiOEP array formed on Au(111) [82]. However, the replacement reaction of the first adlayer of nickel(II) tetraphenylporphyrin (NiTPP) occurred upon adsorption of fullerene molecules as the second layer, suggesting that supramolecular assemblies of C60 and C70 are strongly influenced by the underlying layer of porphyrin, as shown in Fig. 8. The dependency on the underlying porphyrin layer for supramolecular assembly of fullerenes might be tentatively explained by the difference in interac-tion between porphyrin and fullerene. According to the density functional theoretical (DFT) calculation, the interaction energy between ZnTPP and C60 was estimated to be -16 to -18 kcal.mol-1 [140], whereas the interaction energy between ZnP and C60 was -33.9 kcal.mol-1 [141]. Because the adsorption energy of C60 at the Au(111) surface is known to be 16~45 kcal.mol-1 (average is approximately 31 kcal mol-1) [142], this is consistent with the result of the displacement by C60 at NiTPP layer on Au(111) during modification. It has been demonstrated that the stability of the first adlayer is an important factor in the surface design of host-guest selectivity of fullerenes on electrode surfaces [82].

Bonifazi et al. synthesized ZnII porphyrin derivatives 1 and 2 (see Chart 2) and in-vestigated supramolecular assemblies of C60

on 1 and 2 arrays formed on Ag single crystal surfaces in a UHV environment [143]. Such π-conjugated molecules with tunable electronic properties are building blocks for the construction of functional materials with exceptional electrochemical and photo-physical properties. C60 molecules located precisely on top of the 3-cyanophenyl substituents, in a zinc porphyrin derivative 1 array, form a chainlike structure on the zinc porphyrin derivative 1-modified Ag(100) surface have especially exceptional properties as shown in Fig. 9a. The lateral displacements of C60 molecules resulted from the manipulation of the STM tip. The black circle denotes a C60 molecule that vanished during the repositioning operation. As illustrated in the proposed model (Fig. 9b), C60 molecules are located precisely on top of the 3-cyanophenyl substituents in porphyrinatozinc dimer array on Ag(100). In the case of 2, two segregated domains, such as (2√3 × 2√3)R30˚ and a condensed phase composed of 2 were found on Ag(111) with the deposition of 0.14 ML of C60 onto 0.85 ML of 2. The

Fig. 7. High-resolution STM images of TPA-Fe coordination networks on Cu(100); (a) ladder-type structures with two distinct types of nanocavities (marked by A and B), where not all available carboxylate groups are involved in coordination bonding and (b) fully interconnected network with complete 2D reticulation, giving rise to square cavities (marked by C). (c) High-resolution STM image of 2D reticulated Fe-TDA open network with rectangular nanocavities. (d) High-resolution STM image of adsorption of single C60 in Fe-TPA host networks. Inset: C60-monomer accommodation in cavities of type-C network. (e-g) top-view models for C60 adsorption in the cavities encountered in (d) with molecules drawn to scale. Reprinted with permission from reference 125, copyright © 2004 Nature Publishing Group

Chart 2. Chemical formulae of ZnII porphyrin derivatives 1 and 2

Page 11: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

SUPRAMOLECULARLY ASSEMBLED NANOSTRUCTURES 323

unprecedented molecular assembly of 2 and C60 was caused by thermal annealing (453 K). The 2 adlayer recognizes the formation of hybrid 2-C60 assembly with a C60 paired line pattern as the main phase. The C60 molecules were arranged in vertically aligned pairs (intrapair C60-C60 distance of 2.3 nm) with an intermolecular distance of about 6.0 nm, whereas the paired line pattern was repeated every 7.3 nm in the horizontal direction (see Fig. 9c). Thus, supramolecular assembly of C60 also depends on chemical structures in the underlying porphyrin layer or the crystallographic orientation of metal substrates [143].

The control of the supramolecular as-sembly of fullerenes has been extended to another system. Mena-Osteritz et al. synthe-sized macrocyclic oligothiophenes, C[12]T [144], and reported that a self-assembled adlayer of C[12]T on HOPG can be used as a veritable template to epitaxially grow 3D nanoarchitectures with C60 [145]. STM images allowed the investigation of unique 1:1 donor-acceptor complexes comprising a ring-shaped p-type and a spherical n-type semiconductor, under ambient conditions at room temperature. The electronic properties

of the complexes were elucidated by means of scanning tunneling spectroscopy (STS), which showed interesting saturation beha-vior in the I-V curves [145].

We found that supramolecularly assem-bled layers of C60 were formed on both coronene- and perylene-modified Au(111) surfaces [146]. As reported in our previous papers, coronene formed a highly ordered array having a (4 × 4) symmetry on Au(111) [147, 148]. The adlayer structure of the C60 was found to be strongly influenced by the underlying organic layers, suggesting that the latter underlying organic adlayers play an important role in the process of the formation of the C60 molecular adlayer. The all-carbon (C2m-C-C2n type) fullerene dimers such as C60-C-C70 (C131) molecules also formed a stable honeycomb-like array on the coronene-modified Au(111) surface, suggesting that the insertion of one C atom between C60 and C70 cages can alter the adsorption site of each cage. [149]. Epitaxial molecular assemblies for other fullerenes such as C70, C60-C60 dumbbell-dimer (C120), and C60-C70 cross-dimer (C130) were also found on the coronene-modified Au(111) surface. The details of this experiment were reported in a separate paper [150].

Fig. 8. Typical STM images, obtained in 0.1 M HClO4, of C60 array after immersing (a) NiOEP- and (b) NiTPP-modified Au(111) surfaces into C60 benzene solution. (c and d) Schematic illustrations of the supramolecular assembly process of C60 during the modification of NiOEP and NiTPP adlayers, respectively. Reprinted with permission from reference 82, copyright © 2006 The Electrochemical Society of Japan

Fig. 9. (a) STM image (15 × 15 nm2) and (b) the proposed model of the chainlike assembly of C60 on 1-adlayer on Ag(100). (c) STM image (30 × 30 nm2) of the C60 paired line pattern on the 2-adlayer on Ag(111) obtained at 453 K. Reprinted with permission from reference 143, copyright © 2004 Wiley-VCH

Page 12: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

S. YOSHIMOTO AND K. ITAYA324

Dendritic molecules appended with multiple zinc porphyrin units (DPm m; number of ZnII porphyrin moiety) and bipyridine compounds carrying mul-tiple fullerene units (Py2Fn) n; number of C60 units) were synthesized by Aida and coworkers (see Chart 3) [151]. Such segregated arrays of multiple donor and acceptor units using a dendritic scaffold indicate a photoinduced charge separation. Some of the coor-

dination complexes between DPm and Py2Fn were visualized by STM under UHV conditions. For example, DP12 molecules were clearly observed as petal-like patterns with a uniform diameter of 7 nm, as shown in Fig. 10a, when a CHCl3 solution con-taining DP12 and Py2F3 was deposited on a Au(111) surface by pulse injection. High-resolution STM images also exhibited many bright spots at the peri-

phery of DP12 molecules, which are most likely fullerene clusters of Py2F3 (Fig. 10b). The pulse injection method has been already established for observing structures of supramolecular assemblies under UHV conditions at liquid nitrogen temperature. Several UHV-STM studies on covalently linked multi-porphyrins formed on Cu(100) have already been reported [152,153]. An interesting ferrocene-bridged zinc porphyrin macrocycle was synthesized by Shoji et al. [154]. This group succeeded in visualizing the single molecule of a supramolecular coordination assembly on Au(111) in UHV [155].

Control of molecular orientation of functionalized fullerenes. The approach of supramolecular assembly consisting of fullerene and MOEP is more effective for control of molecular orientation of asym-metrical functional fullerenes [50, 156, 157]. For example, open-cage fullerenes are attractive and important molecules for synthesis of a new endohedral fullerene encapsulating molecular hydrogen. Recen-tly, Komatsuʼs group succeeded in prepa-ring H2@C60 and its dimer (H2@C60)2 by a strategy of ʻmolecular surgery ̓in which the fullerene cage is opened and closed after insertion of H2 gas [158, 159]. Subsequently, H2@C70 and 2H2@C70 molecules were also synthesized by the same method [160]. Because the open-cage fullerene derivative 3 (see Chart 4) has two carbonyl moie-ties, the electrochemical redox reaction is expected to be useful as a monitor reaction of the modified electrode surface. Figure 11a shows typical cyclic CVs of a well-defined Au(111) (dotted line) and 3 directly attached on a Au(111) electrode (red solid line) in 0.05 M H2SO4, recorded at a scan rate of 20 mV.s-1. For the Au(111) electro-de directly modified with 3, broad redox peaks were observed at 0 and 0.85 V. In contrast, when the modification by ZnOEP was carried out before the adsorption of 3, a pair of characteristic redox peaks clearly appeared at 0 and 0.85 V during cathodic

Fig. 10. (a) Large-scale and (b) high-resolution STM images of the coordinated complex consisting of DP12 and Py2F3 on Au(111) obtained in UHV. Reprinted with permission from reference 151, copyright © 2006 American Chemical Society

Chart 3. Chemical structures of DP12 and Py2F3

Page 13: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

SUPRAMOLECULARLY ASSEMBLED NANOSTRUCTURES 325

and anodic scans, respectively. This result suggests that 3 was attached to the ZnOEP-modified Au(111) surface, and that carbonyl groups of 3 were oriented toward the solution phase. The CV profile drawn with a solid red line in Fig. 11a is strongly associated with the electrochemical redox reaction of >C=O to >C•-OH (and >C•-OH to >C=O) for two carbonyl groups in each molecule of 3. On the basis of the electronic charge calculated from the reductive peak area, the amount of transferred electronic charge is estimated to be ca. 15.8 μC.cm-2. If two-electron reduction occurs

on the 3/ZnOEP-modified Au(111) electrode, the value corresponds to the surface concentration of (7.9 ± 0.7) × 10-11 mol.cm-2. The electron transfer process is very slow as indicated by the CV profile shown in Fig. 11a. When potential switching was carried out at potentials negative than 0.85 V, no reduction peak at 0 V was observed. The STM image showed a com-pletely disordered structure of 3 directly attached to the surface of Au(111), whereas highly-ordered arrays consisting of bright round spots were found on the ZnOEP-modified Au(111), as shown in Fig. 11b [50, 156]. This result suggests that the formation of a highly ordered adlayer of 3 plays a significant role in the control of molecular orientation. The ad-layer structure of 3 on the ZnOEP-modified Au(111) surface was identical to that of C60 on the Au(111) surface covered with highly ordered ZnOEP arrays.

Furthermore, dependency upon the adlayer struc-tures of ZnOEP was found on reconstructed Au(100)-(hex) and unreconstructed Au(100)-(1 × 1) surfaces [157]. A hexagonally arranged ZnOEP array was formed on a Au(100)-(hex) surface, whereas a rec-tangularly arranged ZnOEP array was found on a Au(100)-(1 × 1) surface. The adlayer structure of ZnOEP was dependent upon the underlying Au atomic

Chart 4. Chemical structures of open-cage C60 and C60Fc derivatives 3 and 4

Fig. 11. Cyclic voltammograms of 3/ZnOEP-adsorbed Au(111), Au(100)-(hex) and Au(100)-(1 × 1) electrodes in 0.05 M H2SO4 recorded at a scan rate of 20 mV.s-1 are shown in (a), (c) and (e), respectively. High-resolution (20 × 20 nm2) STM images of the adlayer of 3 formed on ZnOEP-modified Au(111) for (b), Au(100)-(hex) for (d), and Au(100)-(1 × 1) for (f), all obtained in 0.05 M H2SO4. Also, proposed models are illustrated at the lower part. Inset: a height-shaded view of 3 formed on the ZnOEP-modified Au(111) surface. Reprinted with permissions from reference 156, copyright © 2004 Wiley-VCH and from reference 157, copyright © 2005 American Chemical Society

1 μA.cm-2 1 μA.cm-2 1 μA.cm-2

Page 14: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

S. YOSHIMOTO AND K. ITAYA326

arrangements [157]. The voltammetric responses for the 3 on the ZnOEP-modified Au(100)-(hex) surface were similar to that on the ZnOEP-modified Au(111) surface, as shown in Fig. 11c. On the basis of the electronic charge calculated from the reductive peak area, the amount of transferred electronic charge was estimated to be ca. 16.2 μC.cm-2, which is a similar value obtained at Au(111) electrode. For the 3 on the ZnOEP-modified Au(100)-(1 × 1) electrode, a pair of broad reduction and re-oxidation peaks was observed during the scan (Fig. 11e). Redox peak currents were smaller than those obtained at the reconstructed Au(100)-(hex) surface, suggesting that the molecular orientation of 3 was random on the ZnOEP-modified Au(100)-(1 × 1) surface. This difference in cyclic voltammograms is clearly reflected in STM images for the modified electrodes. Figure 11d shows a typi-cal STM image of an adlayer of 3 formed on ZnOEP-modified Au(100)-(hex) in 0.05 M H2SO4. Highly ordered arrays consisting of bright round spots were observed over the entire surface of Au(100)-(hex). In contrast, a completely disordered structure was ob-served for the adlayer of 3 on the ZnOEP-modified Au(100)-(1 × 1) surface, as shown in Fig. 11f. Although several individual molecules of 3 could be distin-guished under the present conditions, the ZnOEP-modified Au(100)-(1 × 1) surface was largely covered with aggregates of 3, as can be seen. These results indicate that the carbonyl groups of 3 were oriented toward the solution phase on the ZnOEP-modified Au(100)-(hex) surface because of the formation of a 1:1 supramolecular assembly with highly ordered ZnOEP arrays. When a polycrystalline Au electrode, such as a disk or a wire, was used as a substrate, the redox reaction was not evident even on the 3/ZnOEP system. This finding suggests that precise control of underlying ZnOEP adlayer structures with Au ato-mic structure is important to recognize the 3 present on them [157]. The clear enhanced redox peaks in the CV profile at the 3/ZnOEP-modified Au(111) and Au(100)-(hex) electrodes strongly support the conclusion that the orientation of 3 is controlled by the ZnOEP adlayer.

Effect of central metal ions. The supramolecular modification is also effective for control of the elec-tron transfer reaction of ferrocene-linked fullerene (C60Fc), 4 (see Chart 4) [161]. Because ferrocene

is a typical redox species, it is a promising material for an electrochemical switching or sensing device. We investigated OEPs with several metals such as CoOEP, CuOEP, PdOEP, carbon FeClOEP and monoxide-coordinated Ru(CO)OEP as an underlying adsorbed layer. Figure 12a-f shows CVs of 4 on each MOEP-modified Au(111) electrode. Judging from the oxidative peak area, we calculated the electronic charges consumed during the oxidation. The electro-nic charge of redox reaction of ferrocene moiety on each adlayer is summarized in Table 1. The electronic charge of CuOEP consumed by the oxidation reaction of Fc to Fc+ was estimated to be slightly lower than that of CoOEP, ZnOEP, and PdOEP. Furthermore, when FeClOEP and Ru(CO)OEP were used as an underlying adsorbed layer, different CVs were clearly obtained, as shown in Fig. 12e,f. The electrochemical response of the Fc moiety was poor at FeClOEP- and Ru(CO)OEP-modified Au(111) electrode, i.e. the redox peak currents observed around 0.80 V for these two adlayers were much smaller than those of ZnOEP-, CoOEP-, CuOEP- and PdOEP-modified Au(111) electrodes. These results for the FeClOEP- and Ru(CO)OEP-modified Au(111) electrode sug-gest that the surface was not fully covered with 4 molecules. Such metal ion dependence was found in the investigation of affinity between metalloporphy-rin cyclic dimer and fullerene. According to a paper by Aida and coworkers, the association constants, Kassoc values for metalloporphyrin cyclic dimer of Co and Zn ions were greater than 107 M-1, whereas the other three host molecules with other metal ions such as Ni, Cu, and Ag were inferior to the above host molecules (Kassoc < 107 M-1) [162]. Recently, DFT calculations for other central metal ions were also reported by Basiuk [141]. DFT calculations for supra-molecular interactions between metalloporphyrin and C60 showed that the interaction energy is strongly influenced by central metal ions. The decrease in the HOMO-LUMO gap energy of porphyrin framework is estimated to be found by supramolecular assembly of C60, as shown in Fig. 13. Although the electronic property of MOEP layers might also be affected by a gold surface, the difference in the HOMO-LUMO gap energy depending upon the central metal ions might be explained by the difference in electronic charge (or peak current) from the redox reaction of the Fc moiety.

Table 1. Electronic charge calculated from the oxidative peak area and C60Fc adlayers based on the STM observation for each MOEP layer

Compound ZnOEP CoOEP CuOEP PdOEP FeClOEP Ru(CO)OEP

Q (μC.cm-2) 8.0 6.1 4.8 6.5 — —

Orientation of C60Fc ordered ordered ordered ordered disordered disordered

Page 15: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

SUPRAMOLECULARLY ASSEMBLED NANOSTRUCTURES 327

To understand the details of the interfacial phenomena, STM measurements were carried out in 0.1 M HClO4 for each case [161]. Hexagonal molecular packing arrangements were clearly obser-ved by the same procedure for six different metal-coordinated OEPs (see Fig. 12g-l). Figure 12m-r shows typical STM images of the 4 arrays formed on ZnOEP-, CoOEP-, CuOEP-, PdOEP-, FeClOEP- and Ru(CO)OEP-modified Au(111) surfaces, respec-tively. As can be seen in Fig. 12g,p, 4 molecules were hexagonally arranged. Adlattices of 4 on the ZnOEP, CoOEP, CuOEP and PdOEP adlayers were almost identical to C60 on the ZnOEP adlayer. In contrast, Figure 12q,r shows typical large-scale STM images of 4 array on the FeClOEP- and Ru(CO)OEP-modi-fied Au(111) surfaces. Although some ordered areas are visible, the entire surface was rough. Thus, the axial coordination of Cl and CO is likely to signifi-cantly affect the fullerene and porphyrin supra-molecular assembly on metal surfaces, indicating that the coordination ligand in OEP faces the solution phase, in contrast with the Au(111) surface. The effect of the central metal ion in OEP on the formation of supramolecular assemblies with the 4 molecule was clearly demonstrated in the CV profiles and STM images. The same tendency was also found at supramolecular assembled layers with 3 [163]. Thus, based on the electrochemical results, MOEP adlayer

plays an important role in not only suppression of the oxidative desorption of C60 but also the control of orientation of the C60 derivatives, because the fullerenes can be trapped in the “cup” consisting of ethyl moieties in the OEP scaffold.

Nanoapplications. Fullerenes are often used as a model compound for the manipulation (guest mo-lecule) of surface-supported supramolecular two-dimensional arrays both in UHV and in solution. Hecklʼs group found that trimesic acid (TMA) forms two different structures on HOPG, a honeycomb lattice in the connection between the hexagons (cyclic hexamers) and a flower-like motif through hydrogen bonding between three TMA molecules between neighboring rings in UHV [164] and at the solid-liquid interface [165]. According to their study, it was found that the honeycomb lattice structures, whether hexagons or flower-like motif, were strongly influen-ced by organic solvents such as butyric acid, pentanoic acid, hexanoic acid, heptanoic acid, octanoic acid and nonanoic acid [165]. Adlayer structures of TMA on HOPG surface depended on the solvent. Such cyclic hexamers stabilized by hydrogen bonding among TMA molecules could act as a host network structure for guest molecules such as coronene [166] and C60 molecules [167]. In manipulation experiments with C60 molecules as a guest in the chicken wire structure,

Fig. 12. Cyclic voltammograms recorded at scan rate of 50 mV.s-1 of 4 on (a) ZnOEP-, (b) CoOEP-, (c) CuOEP-, (d) PdOEP-, (e) FeClOEP- and (f) Ru(CO)OEP-modified Au(111) (red solid line) in pure 0.1 M HClO4. The dotted lines show voltammograms of the underlying MOEP-modified Au(111) electrode. High-resolution STM images of (g) ZnOEP, (h) CoOEP, (i) CuOEP, (j) PdOEP, (k) FeClOEP and (l) Ru(CO)OEP underlying layers on Au(111) surface, and of 4 arrays on (m) ZnOEP-, (n) CoOEP-, (o) CuOEP-, (p) PdOEP-, (q) FeClOEP- and (r) Ru(CO)OEP-modified Au(111) obtained at 0.75 V vs. RHE in 0.1 M HClO4. Reprinted with permission from reference 161 for figures (a-c), (e), (h), (i), (k), (m-o), and (q), copyright © 2004 American Chemical Society. Note that CVs and STM images for PdOEP and Ru(CO)OEP are unpublished results. The scales of individual STM images were 15 × 15 nm2 for (g) -(k), 10 × 10 nm2 for (l), and 30 × 30 nm2 for (m)-(r), respectively

Page 16: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

S. YOSHIMOTO AND K. ITAYA328

the transfer from one cavity to an adjacent cavity was demonstrated.

Spillmann et al. reported that 2 molecules domi-nantly formed a hexagonal pattern by trimer forma-tion, when 0.5-0.7 MLs of porphyrin 2 was deposited on Ag(111) [168]. Since the 3-cyanophenyl residues and the central porphyrin core hardly contribute to the tunneling current, each single porphyrin molecule appears as two bright lobes separated by 1.20 nm, which is consistent with the intramolecular distan-ce between the 3,5-di(tert-butyl)phenyl substituents (center-to-center distance: 1.26 nm) in the crystal structure of 2. As reported by Yokoyama et al., it is suggested that the molecular assembly is driven by both van der Waals and dipole-dipole interactions involving the polar 3-cyanophenyl residues [123]. After further deposition of C60 molecules, a lateral displacement of the C60 molecules on the porphyrin

network was found in the time-dependent STM images, as shown in Fig. 14a-c. To minimize the in-fluence of the tip on the hopping process, all STM images were recorded under the condition of a high tunneling gap resistance (>100 GΩ). Driven by thermal fluctuations, single C60 ad-molecules displace to neighboring pores as time proceeds. Fur-thermore, the conformational motion of the 3,5-di-(tert-butyl)phenyl moieties in the porous network and self-repairing of a defect in the porous structure of 2 (dashed ellipse) were also found in those STM images. The same group subsequently reported the difference in mobility of single fullerene molecules between C60 and C70 [169]. The hopping rate of C60 was 1 × 10-3 s-1 at coverage of 0.06, whereas that of C70 was 9 × 10-2 s-1 at coverage of 0.01. The hopping rate also depends upon the coverage of fullerene on the porphyrin network.

Fig. 13. Density functional theory (DFT) calculations (HOMO-LUMO) of each metalloporphyrin and supramolecular assembly with C60. Reprinted with permission from reference 141, copyright © 2005 American Chemical Society

Page 17: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

SUPRAMOLECULARLY ASSEMBLED NANOSTRUCTURES 329

Shirai et al. synthesized unique and attractive molecules, the so-called “nanocars” molecules con-sisting of an oligo(phenylenethylene) “chassis” supporting four C60-derived “wheels” (see Fig. 15a,b) [170, 171]. The motion of the “nanocar” depended upon the temperature. The “nanocar” remained effec-

tively stationary on the surface at room temperature up to approximately 170 °C. The pivoting and translational motions of 5 were observed on a Au(111) terrace at approximately 200 °C, as seen in a se-quence of STM images in Fig. 15c-f. The translational motion that occurred between pivoting was perpendicular to the axles, illustrating a directional preference relative to the molecular orientation. The observed movement of the nanocars indicated a new type of wheel-like rolling motion based on fullerene, as opposed to stick-slip or sliding translation. These molecules which include ʻnanotrains ̓ and ʻnanotrucks ̓ are promi-sing materials for assisted small molecule transport across surfaces. Further investi-gations are expected for the development of attractive molecular motions.

SUMMARY AND OUTLOOKThis minireview primarily reported self-

organization and supramolecular assembly of porphyrins, phthalocyanines, and fulle-renes at surfaces using STM. The under-standing of both intermolecular and mole-cule-substrate interactions in two-dimen-sionally self-organized films at the molecu-lar level in UHV as well as in solution has been advanced considerably by the use of STM techniques in the past decade. The ʻbottom-up ̓ strategy is an attractive and promising approach for construction of nanoarchitectures. The understanding and the construction of the surface-supported supramolecular assembly have provided knowledge necessary for new surface de-sign and patterning using characteristic molecular assemblies by non-covalent in-termolecular interactions, such as dipole-dipole interactions, hydrogen bonding, elec-trostatic interactions, metal-ligand coordi-nation and π-π interactions. To produce new functional materials and molecular devices based on the above knowledge, it is neces-sary to continue further exploration of the nanostructures at surfaces and the design of attractive molecules by organic synthesis in the future. However, as a next step, it is important to demonstrate a special function such as catalysis, photo-induced electron

transfer, and conductivity (or hole-mobility) from the formation of nanostructures. The demonstration of the special functions by the formation of nanostructures must provide a “breakthrough” to the next stage of the ʻbottom up ̓strategy.

Fig. 14. Time-dependent STM images (25 × 25 nm2) for (a, c, and d) obtained at 298 K (each image was taken in every 62 s). The dotted circle indicates the open portion after single C60 ad-molecule displaces to neighboring pores by thermal fluctuations. (b) Proposed molecular model of a C60 molecule hosted inside a supramolecular porphyrin-based pore. Reprinted with permission from reference 168, copyright © 2006 Wiley-VCH

Fig. 15. (a) High-resolution STM image and (b) model structure of ʻnanocar ̓5. (c-f) Time-dependent STM images for the rolling of 5 on Au(111) in UHV. Reprinted with permission from reference 170, copyright © 2005 American Chemical Society

Page 18: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

S. YOSHIMOTO AND K. ITAYA330

Acknowledgements

The authors would like to express sincerest thanks to Prof. Osamu Ito for his invaluable com-ments, suggestions and discussions. His constant encouragement conferred a challenging spirit for SY during the time that SY was working at Tohoku University (April 2000 to March 2005). This work was supported in part by the Ministry of Education, Culture, Sports, Science and Technology, through a Grant-in-Aid for Young Scientists (B) (No. 18750132) and by the Center of Excellence (COE) Project, Giant Molecules and Complex Systems, 2007.

REFERENCES 1. Balzani V. (Ed) Electron Transfer in Chemistry

Vol. 3, Wiley-VCH: New York, 2001. 2. Fukuzumi S. Org. Biomol. Chem. 2003; 1: 609-

620. 3. Guldi DM. Chem. Soc. Rev. 2002; 31: 22-36. 4. Imahori H, Mori Y and Matano Y. J. Photochem.

Photobiol. C 2003; 4: 51-83. 5. Imahori H. Org. Biomol. Chem. 2004; 2: 1425-

1433. 6. Imahori H and Fukuzumi S. Adv. Funct. Mater.

2004; 14: 525-536. 7. El-Khouly ME, Ito O, Smith PM and DʼSouza

F. J. Photochem. Photobiol. C 2004; 5: 79-104. 8. Hasobe T, Fukuzumi S and Kamart PV. Interface

2006; 15(2): 47-51. 9. Collman JP, Gagne RR, Halbert TR, Marchon

JC and Reed CA. J. Am. Chem. Soc. 1973; 95: 7868-7870.

10. Collman JP, Gagne RR, Reed CA, Halbert TR, Lang G and Robinson WT. J. Am. Chem. Soc. 1975; 97: 1427-1439.

11. Collman JP and Fu L. Acc. Chem. Res. 1999; 32: 455-463.

12. Collman JP, Boulatov R, Sunderland CJ and Fu L. Chem. Rev. 2004; 104: 561-588.

13. Collman JP, Marrocco M, Denisevich P, Konai Y, Koval C and Anson FC. J. Electroanal. Chem., Interfacial Electrochem. 1979; 101: 117-122.

14. Collman JP, Denisevich P, Konai Y, Marrocco M, Koval C and Anson FC. J. Am. Chem. Soc. 1980; 102: 6027-6036.

15. Yeager E. Electrochim. Acta 1984; 29: 1527-1537, and references therein.

16. Collman JP, Wagenknecht PS and Hutchison JE. Angew. Chem., Int. Ed. 1994; 33: 1537-1554.

17. Imahori H, Hagiwara K, Aoki M, Akiyama T, Taniguchi S, Okada T, Shirakawa M and Sakata Y. J. Am. Chem. Soc. 1996; 118: 11771-11782.

18. Kuciauskas D, Lin S, Seely GR, Moore AL, Moore TA, Gust D, Drovetskaya T, Reed CA and Boyd PDW. J. Phys. Chem. 1996; 100:

15926-15932.19. Sun Y, Drovetskaya T, Bolskar RD, Bau R,

Boyd PDW and Reed CA. J. Org. Chem. 1997; 62: 3642-3649.

20. Thilgen C and F Diederich F. Chem. Rev. 2006; 106: 5049-5135.

21. Tashiro K, Aida T, Zheng JY, Kinbara K, Saigo K, Sakamoto S and Yamaguchi K. J. Am. Chem. Soc. 1999; 121: 9477-9478.

22. Tashiro K and Aida T. Chem. Soc. Rev. 2007; 36: 189-197.

23. Sun D, Tham FS, Reed, CA, Chaker L, Burgess M and Boyd PDW. J. Am. Chem. Soc. 2000; 122: 10704-10705.

24. Gust D, Moore TA and Moore AL. Acc. Chem. Res. 2001; 34: 40-48.

25. Boyd PDW and Reed CA. Acc. Chem. Res. 2005; 38: 235-242.

26. Lehn JM. Supramolecular Chemistry Wiley-VCH: Weinheim, 1995.

27. Lehn JM. Science 2002; 295: 2400-2403.28. Lehn JM. Chem. Soc. Rev. 2007; 36: 151-160.29. Choi IS, Bowden N and Whitesides GM. Angew.

Chem., Int. Ed. 1999; 38: 3078-3081.30. Li G, Fudickar W, Skupin M, Klyszcz A, Draeger

C, Lauer M, Fuhrhop JH. Angew. Chem., Int. Ed. 2002; 41: 1828-1852.

31. Ruben M, Rojo J, Romero-Salguero FJ, Uppadine LH, and Lehn JM. Angew. Chem., Int. Ed. 2004; 43: 3644-3662.

32. Diederich F. Angew. Chem., Int. Ed. 2007; 46: 68-69.

33. Bonifazi D, Enger O and Diederich F. Chem. Soc. Rev. 2007; 36: 390-414.

34. van Hameren R, Schön P, van Buul AM, Hoogboom J, Lazarenko SV, Gerritsen JW, Engelkamp H, Christianen PCM, Heus HA, Maan JC, Rasing T, Speller S, Rowan AE, Elemans JAAW and Nolte RJM. Science, 2006; 314: 1433-1436.

35. Hill JP, Jin W, Kosaka A, Fukushima T, Ichihara H, Shimomura T, Ito K, Hashizume T, Ishii N and Aida T. Science 2004; 304: 1481-1483.

36. Rothemund PWK, Nature, 2006; 440: 297-302.37. Yan H, Park SH, Finkelstein G, Reif JH and

LaBean TH. Science 2003; 301: 1882-1884.38. Park SH, Pistol C, Ahn SJ, Reif JH, Lebeck AR,

Dwyer C and LaBean TH. Angew. Chem., Int. Ed. 2006; 45: 735-739.

39. Lin C, Liu Y, Rinker S and Yan H. ChemPhysChem 2006; 7: 1641-1647.

40. Lin C, Liu Y and Yan H. Nano Lett. 2007; 7: 507-512.

41. Chiang S. Chem. Rev. 1997; 97: 1083-1096.42. Poirier GE. Chem. Rev. 1997; 97: 1117-1127.43. De Feyter S and De Schryver FC. Chem. Soc.

Rev. 2003; 32: 139-150.

Page 19: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

SUPRAMOLECULARLY ASSEMBLED NANOSTRUCTURES 331

44. Barth JV, Costantini G and Kern K. Nature 2005; 437: 671-679.

45. Gewirth AA and Niece BK. Chem. Rev. 1997; 97: 1129-1162.

46. Itaya K. Prog. Surf. Sci. 1998; 58: 121-248.47. Kolb DM. Angew. Chem., Int. Ed. 2001; 40:

1162-1181.48. Magnussen OM, Chem. Rev. 2002; 102: 672-

725.49. De Feyter S and De Schryver FC. J. Phys. Chem.

B 2005; 109: 4290-4302.50. Yoshimoto S. Bull. Chem. Soc. Jpn. 2006; 79:

1167-1190.51. Wan LJ. Acc. Chem. Res. 2006; 39: 334-342.52. Lippel PH, Wilson RJ, Miller MD, Wöll C and

Chiang S. Phys. Rev. Lett. 1989; 62: 171-174.53. Jung TA, Schlittler RR, Gimzewski JK, Tang H

and Joachim C. Science 1996; 271: 181-184.54. Jung TA, Schlittler RR and Gimzewski JK.

Nature 1997; 386: 696-968.55. Lu X, Hipps KW, Wang XD and Mazur U. J.

Am. Chem. Soc. 1996; 118: 7197-7102.56. Hipps KW, Lu X, Wang XD and Mazur U. J.

Phys. Chem. 1996; 100: 11207-11210.57. Lu X and Hipps KW. J. Phys. Chem. B 1997;

101: 5391-5396.58. Barlow DE and Hipps KW. J. Phys. Chem. B

2000; 104: 5993-6000.59. Scudiero L, Barlow DE and Hipps KW. J. Phys.

Chem. B 2000; 104: 11899-11905.60. Scudiero L, Barlow DE, Mazur U and Hipps

KW. J. Am. Chem. Soc. 2001; 123: 4073-4080.61. Kunitake M, Batina N and Itaya K. Langmuir

1995; 11: 2337-2340.62. Batina N, Kunitake M and Itaya K. J. Electroanal.

Chem. 1996; 405: 245-250.63. Kunitake M, Akiba U, Batina N and Itaya K.

Langmuir 1997; 13: 1607-1615.64. Ogaki K, Batina N, Kunitake M and Itaya K. J.

Phys. Chem. 1996; 100: 7185-7890.65. Sashikata K, Sugata T, Sugimasa M and Itaya K.

Langmuir 1998; 14: 2896-2902.66. Wan LJ, Shundo S, Inukai J and Itaya K.

Langmuir 2000; 16: 2164-2168.67. He Y, Ye T and Borguet E. J. Am. Chem. Soc.

2002; 124: 11964-11970.68. Ye T, He Y and Borguet E. J. Phys. Chem. B

2006; 110: 6141-6147.69. Tao NJ, Cardenas G, Cunha F and Shi Z.

Langmuir 1995; 11: 4445-4448.70. Tao NJ. Phys. Rev. Lett. 1996; 76: 4066-4069.71. Yoshimoto S, Inukai J, Tada A, Abe T, Morimoto

T, Osuka A, Furuta H and Itaya K. J. Phys. Chem. B 2004; 108: 1948-1954.

72. Yoshimoto S, Tada A, Suto K, Narita R and Itaya K. Langmuir 2003; 19: 672-677.

73. Yoshimoto S, Sato K, Sugawara S, Chen Y, Ito

O, Sawaguchi T, Niwa O and Itaya K. Langmuir 2007; 23: 809-816.

74. Yoshimoto S, Tada A, Suto K and Itaya K. J. Phys. Chem. B 2003; 107: 5836-5843.

75. Scudiero L, Barlow DE and Hipps KW. J. Phys. Chem. B 2002; 106: 996-1003.

76. Yoshimoto S, Yokoo N, Fukuda T, Kobayashi N and Itaya K. Chem. Commun. 2006; 500-502.

77. Yoshimoto S, Suto K, Itaya K and Kobayashi N. Chem. Commun. 2003; 2174-2175.

78. Yoshimoto S, Suto K, Tada A, Kobayashi N and Itaya K. J. Am. Chem. Soc. 2004; 126: 8020-8027.

79. Yoshimoto S, Tada A, Suto K, Yau SL and Itaya K. Langmuir 2004; 20: 3159-3165.

80. Yoshimoto S, Tsutsumi E, Honda Y, Ito O and Itaya K. Chem. Lett. 2004; 33: 914-915.

81. Yoshimoto S, Tsutsumi E, Suto K, Honda Y and Itaya K. Chem. Phys. 2005; 319: 147-158.

82. Yoshimoto S, Sugawara S and Itaya K. Electrochemistry 2006; 74: 175-178.

83. Gottfried JM, Flechtner K, Kretschmann A, Lukasczyk T and Steinrück HP. J. Am. Chem. Soc. 2006; 128: 5644-5645.

84. Kretschmann A, Walz MM, Flechtner K, Steinrück HP and Gottfried JM. Chem. Commun. 2007; 568-570.

85. Auwärter W, Weber-Bargioni A, Brink S, Riemann A, Schiffrin A, Ruben M and Barth JV. ChemPhysChem 2007; 8: 250-254.

86. Buchner F, Schwald V, Comanici K, Steinrück HP and Marbach H. ChemPhysChem 2007; 8: 241-243.

87. Hunter CA, Mesh MN and Sanders JKM. J. Am. Chem. Soc. 1990; 112: 5773-5780.

88. Anderson HL, Hunter CA, Mesh MN and Sanders JKM. J. Am. Chem. Soc. 1990; 112: 5780-5789.

89. Anderson S, Anderson HL and Sanders JKM. Acc. Chem. Res. 1993; 26: 469-475.

90. Tsuda A and Osuka A. Science 2001; 293: 79-82.

91. Hwang IW, Kamada T, Ahn TK, Ko DM, Nakamura T, Tsuda A, Osuka A and Kim D. J. Am. Chem. Soc. 2004; 126: 16187-16198.

92. van Gerven PCM, Elemans JAAW, Gerritsen JW, Speller S, Nolte RJM and Rowan AE. Chem. Commun. 2005; 3535-3537.

93. Deiters E, Bulach V and Hosseini MW. Chem. Commun. 2005; 3906-3908.

94. Hosseini MW. Chem. Commun. 2005; 5825-5829.

95. Hipps KW, Scudiero L, Barlow DE and Cooke, Jr MP. J. Am. Chem. Soc. 2002; 124: 2126-2127.

96. Scudiero L, Hipps KW and Barlow DE. J. Phys. Chem. B 2003; 107: 2903-2909.

Page 20: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

S. YOSHIMOTO AND K. ITAYA332

97. Barlow DE, Scudiero L and Hipps KW. Langmuir 2004; 20: 4413-4421.

98. Suto K, Yoshimoto S and Itaya K. J. Am. Chem. Soc. 2003; 125: 14976-14977.

99. Suto K, Yoshimoto S and Itaya K. Langmuir 2006; 22: 10766-10776.

100. Yoshimoto S, Higa N and Itaya K. J. Am. Chem. Soc. 2004; 126: 8540-8545.

101. Yang ZY, Gan LH, Lei SB, Wan LJ, Wang C and Jiang JZ. J. Phys. Chem. B 2005; 109: 19859-19865.

102. Klymchenko AS, Sleven J, Binnemans K and De Feyter S. Langmuir 2006; 22: 723-728.

103. Takami T, Arnold DP, Fuchs AV, Will GD, Goh R, Waclawik ER, Bell JM, Weiss PS, Sugiura Ki, Liu W and Jiang JZ. J. Phys. Chem. B 2006; 110: 1661-1664.

104. Ye T, Takami T, Wang R, Jiang JZ and Weiss PS. J. Am. Chem. Soc. 2006; 128: 10984-10985.

105. Takami T, Ye T, Arnold DP, Sugiura K-i, Wang R, Jiang JZ and Weiss PS. J. Phys. Chem. C 2007; 111: 2077-2080.

106. Otsuki J, Kawaguchi S, Yamakawa T, Asakawa M and Miyake K. Langmuir 2006; 22: 5708-5715.

107. Ma H, Ou Yang LY, Pan N, Yau SL, Jiang JZ and Itaya K. Langmuir 2006; 22: 2105-2111.

108. Kottas GS, Clarke LI, Horinek D and Michl J. Chem. Rev. 2005; 105: 1281-1376.

109. Liu Z, Yasseri AA, Lindsey JS and Bocian DF. Science 2003; 302: 1543-1545.

110. Takeuchi M, Imada T, Shinkai S. Angew. Chem., Int. Ed. 1998; 37: 2096-2099.

111. Sugasaki A, Ikeda M, Takeuchi M and Shinkai S. Angew. Chem., Int. Ed. 2000; 39: 3839-3842.

112. Shinkai S, Ikeda M, Sugasaki A and Takeuchi M. Acc. Chem. Res. 2001; 34: 494-503.

113. Tashiro K, Konishi K and Aida T. J. Am. Chem. Soc. 2000; 122: 7921-7926.

114. Yoshimoto S, Sawaguchi T, Su W, Jiang JZ and Kobayashi N. Angew. Chem., Int. Ed. 2007; 46: 1071-1074.

115. Theobald JA, Oxtoby NS, Phillips MA, Champness NR and Beton PH. Nature 2003; 424: 1029-1031.

116. Theobald JA, Oxtoby NS, Champness NR, Beton PH and Dennis TJS. Langmuir 2005; 21: 2038-2041.

117. Perdigão LMA, Perkins EW, Ma J, Staniec PA, Rogers BL, Champness NR and Beton PH. J. Phys. Chem. B 2006; 110: 12539-12542.

118. Xu B, Tao CG, Cullen WG, Reutt-Robey JE and Williams ED. Nano Lett. 2005; 5: 2207-2211.

119. Xu B, Tao CG, Williams ED and Reutt-Robey JE. J. Am. Chem. Soc. 2006; 128: 8493-8499.

120. Lei SB, Wang C, Yin SX, Wang HN, Wi F, Liu HW, Xu B, Wan LJ and Bai CL. J. Phys. Chem.

B 2001; 105: 10838-10841.121. Otsuki J, Nagamine E, Kondo T, Iwasaki K,

Asakawa M and Miyake K. J. Am. Chem. Soc. 2005; 127: 10400-10405.

122. Yokoyama T, Kamikado T, Yokoyama S and Mashiko S. J. Chem. Phys. 2004; 121: 11993-11997.

123. Yokoyama T, Yokoyama S, Kamikado T, Okuno Y and Mashiko S. Nature 2001; 413: 619-621.

124. Nishiyama F, Yokoyama T, Kamikado T, Yokoyama S, Mashiko S, Sakaguchi K and Kikuchi K. Adv. Mater. 2007; 19: 117-120.

125. Stepanow S, Lingenfelder M, Dmitriev A, Spillmann H, Delvigne E, Lin N, Deng X, Cai C, Barth JV and Kern K. Nature Mater. 2004; 3: 229-233.

126. Lingenfelder MA, Spillmann H, Dmitriev A, Stepanow S, Lin N, Barth JV and Kern K. Chem.-Eur. J. 2004; 10: 1913-1919.

127. Spillmann H, Dmitriev A, Lin N, Messina P, Barth JV and Kern K. J. Am. Chem. Soc. 2003; 125: 10725-10728.

128. Dmitriev A, Spillmann H, Lingenfelder M, Lin N, Barth JV and Kern K. Langmuir 2004; 20: 4799-4801.

129. Stepanow S, Lin N, Barth, JV and Kern K. Chem. Commun. 2006; 2153-2155.

130. Clair S, Pons S, Brune H, Kern K and Barth JV. Angew. Chem., Int. Ed. 2005; 44: 7294-7297.

131. Clair S, Pons S, Dmitriev A, Fabris S, Baroni S, Brune H, Kern K and Barth JV. J. Phys. Chem. B 2006; 110: 5627-5632.

132. Stepanow S, Lin N, Payer D, Schlickum U, Klappenberger F, Zoppellaro G, Ruben M, Brune H, Barth JV and Kern K. Angew. Chem., Int. Ed. 2007; 46: 710-713.

133. Ishii T, Aizawa N, Kanehama R, Yamashita M, Sugiura K and Miyasaka H. Coord. Chem. Rev. 2002; 226: 113-121.

134. Stöhr M, Wagner T, Gabriel M, Weyers B and Möller R. Adv. Funct. Mater. 2001; 11: 175-178.

135. Altman EI and Colton RJ. Surf. Sci. 1992; 279: 49-67.

136. Altman EI and Colton RJ. Surf. Sci. 1993; 295: 13-33.

137. Altman EI and Colton RJ. Phys. Rev. B 1993; 48: 18244-18249.

138. Uemura S, Ohira A, Sakata M, Taniguchi I, Kunitake M and Hirayama C. Langmuir 2001; 17: 5-7.

139. Yoshimoto S, Narita R, Tsutsumi E, Matsumoto M, Itaya K, Ito O, Fujiwara K, Murata Y and Komatsu K. Langmuir 2002; 18: 8518-8522.

140. Wang YB and Lin Z. J. Am. Chem. Soc. 2003; 125, 6072-6073.

141. Basiuk VA. J. Phys. Chem. A 2005; 109: 3704-

Page 21: Advances in supramolecularly assembled nanostructures of ...atom/japanese/paper/yoshi.pdf · Supramolecular assemblies based on non-covalent interactions have been explored ... such

Copyright © 2007 Society of Porphyrins & Phthalocyanines J. Porphyrins Phthalocyanines 2007; 11: 313-333

SUPRAMOLECULARLY ASSEMBLED NANOSTRUCTURES 333

3710.142. Pérez-Jiménez ÁJ, Palacios JJ, Louis E,

SanFabián E and Vergés JA. ChemPhysChem 2003; 4: 388-392.

143. Bonifazi D, Spillmann H, Kiebele A, de Wild M, Seiler P, Cheng F, Jung T and Diederich F. Angew. Chem., Int. Ed. 2004; 43: 4759-4763.

144. Krömer J, Rios-Carreras I, Fuhrmann G, Musch C, Wunderlin M, Debaerdemaeker T, Mena-Osteritz E and Bäuerle P. Angew. Chem., Int. Ed. 2000; 39: 3481-3486.

145. Mena-Osteritz E and Bäuerle P. Adv. Mater. 2006; 18: 447-451.

146. Yoshimoto S, Tsutsumi E, Fujii O, Narita R and Itaya K. Chem. Commun. 2005; 1188-1190.

147. Yoshimoto S, Narita R and Itaya K. Chem. Lett. 2002; 356-357.

148. Yoshimoto S, Narita R, Wakisaka M and Itaya K. J. Electroanal. Chem. 2002; 532: 331-335.

149. Zhao Y, Chen Z, Yuan H, Gao X, Qu L, Chai Z, Xing G, Yoshimoto S, Tsutsumi E and Itaya K. J. Am. Chem. Soc. 2004; 126: 11134-11135.

150. Yoshimoto S, Tsutsumi E, Narita R, Murata Y, Murata M, Fujiwara K, Komatsu K, Ito O and Itaya K. J. Am. Chem. Soc. 2007; 129: 4366-4376.

151. Li WS, Kim KS, Jiang DL, Tanaka H, Kawai T, Kwon JH, Kim D and Aida T. J. Am. Chem. Soc. 2006; 128: 10527-10532.

152. Sugiura K, Tanaka H, Matsumoto T, Kawai T and Sakata Y. Chem. Lett. 1999; 1193-1194.

153. Kato A, Sugiura Ki, Miyasaka H, Tanaka H, Kawai T, Sugimoto M and Yamashita M. Chem. Lett. 2004; 33: 578-579.

154. Shoji O, Okada S, Satake A and Kobuke Y. J. Am. Chem. Soc. 2005; 127: 2201-2210.

155. Shoji O, Tanaka H, Kawai T and Kobuke Y. J. Am. Chem. Soc. 2005; 127: 8598-8599.

156. Yoshimoto S, Tsutsumi E, Honda Y, Murata Y, Murata M, Komatsu K, Ito O and Itaya K. Angew. Chem., Int. Ed. 2004; 43: 3044-3047.

157. Yoshimoto S, Honda Y, Murata Y, Murata M, Komatsu K, Ito O and Itaya K. J. Phys. Chem. B 2005; 109: 8547-8550.

158. Komatsu K, M. Murata M and Murata Y. Science 2005; 307: 238-240.

159. Komatsu K and Murata Y. Chem. Lett. 2005; 34, 886-891.

160. Murata M, Murata Y and Komatsu K. J. Am. Chem. Soc. 2006; 128: 8024-8033.

161. Yoshimoto S, Saito A, Tsutsumi E, DʼSouza F, Ito O and Itaya K. Langmuir 2004; 20, 11046-11052.

162. Zheng JY, Tashiro K, Hirabayashi Y, Kinbara K, Saigo K, Aida T, Sakamoto S and Yamaguchi K. Angew. Chem., Int. Ed. 2001; 40: 1858-1861.

163. Yoshimoto S, Saito A and Itaya K. unpublished results.

164. Griessl S, Lackinger M, Edelwirth M, Hietschold M and Heckl WM. Single Mol. 2002; 3: 25-31.

165. Lackinger M, Griessl S, Heckl WM, Hietschold M and Flynn GW. Langmuir 2005; 21: 4984-4988.

166. Griessl SJH, Lackinger M, Jamitzky F, Markert T, Hietschold M and Heckl WM. Langmuir 2004; 20: 9403-9407.

167. Griessl SJH, Lackinger M, Jamitzky F, Markert T, Hietschold M and Heckl WM. J. Phys. Chem. B 2004; 108: 11556-11560.

168. Spillmann H, Kiebele A, Stöhr M, Jung TA, Bonifazi D, Cheng F and Diederich F. Adv. Mater. 2006; 18: 275-279.

169. Kiebele A, Bonifazi D, Cheng F, Stöhr M, Diederich F, Jung T and Spillmann, H. ChemPhysChem 2006; 7: 1462-1470.

170. Shirai Y, Osgood AJ, Zhao Y, Kelly KF and Tour JM. Nano Lett. 2005; 5: 2330-2334.

171. Shirai Y, Osgood AJ, Zhao Y, Yao Y, Saudan L, Yang H, Yu-Hung C, Alemany LB, Sasaki T, Morin JF, Guerrero JM, Kelly KF and Tour JM, J. Am. Chem. Soc. 2006; 128: 4854-4864.


Recommended