Next Article in Journal
Selective Cannabinoid 2 Receptor Agonists as Potential Therapeutic Drugs for the Treatment of Endotoxin-Induced Uveitis
Next Article in Special Issue
Engineering 3D Graphene-Based Materials: State of the Art and Perspectives
Previous Article in Journal
Phytochemical Investigation of Tradescantia Albiflora and Anti-Inflammatory Butenolide Derivatives
Previous Article in Special Issue
Graphene Oxide Tablets for Sample Preparation of Drugs in Biological Fluids: Determination of Omeprazole in Human Saliva for Liquid Chromatography Tandem Mass Spectrometry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Intramolecular Energy- and Electron-Transfer of 3D-Conformeric Tris(fluorenyl-[60]fullerenylfluorene) Derivatives

1
Department of Chemistry, University of Massachusetts Lowell, Lowell, MA 01854, USA
2
Functional Materials Division, AFRL/RXA, Air Force Research Laboratory, Wright-Patterson Air Force Base, Dayton, OH 45433, USA
*
Author to whom correspondence should be addressed.
Molecules 2019, 24(18), 3337; https://doi.org/10.3390/molecules24183337
Submission received: 3 August 2019 / Revised: 30 August 2019 / Accepted: 10 September 2019 / Published: 13 September 2019
(This article belongs to the Special Issue Materials Chemistry of Fullerenes, Graphenes and Carbon Nanotubes)

Abstract

:
New 3D conformers were synthesized to show a nanomolecular configuration with geometrically branched 2-diphenylaminofluorene (DPAF-C2M) chromophores using a symmetrical 1,3,5-triaminobenzene ring as the center core for the connection of three fused DPAF-C2M moieties. The design led to a class of cis-cup-tris[(DPAF-C2M)-C60(>DPAF-C9)] 3D conformers with three bisadduct-analogous <C60> cages per nanomolecule facing at the same side of the geometrical molecular cis-cup-shape structure. A sequential synthetic route was described to afford this 3D configurated conformer in a high yield with various spectroscopic characterizations. In principle, a nanostructure with a non-coplanar 3D configuration in design should minimize the direct contact or π-stacking of fluorene rings with each other during molecular packing to the formation of fullerosome array. It may also prevent the self-quenching effect of its photoexcited states in solids. Photophysical properties of this cis-cup-conformer were also investigated.

Graphical Abstract

1. Introduction

Photoinduced intramolecular energy and electron transfer phenomena in organo [60]fullerene derivatives having a covalent molecular composition of both an electron donor and a [60]fullerenyl or nanocarbon acceptor components were demonstrated over a number of years [1,2,3]. The energy process may involve the facile triplet state of fullerene and other chromophores [4,5,6,7,8]. This type of nanomolecular system was used in many technological applications [9,10], including photovoltaic devices [11,12], sensors and switches [13], and photodynamic therapy [14,15]. Fullerene-based nanostructures with multiple C60 cages [16] in the structure were found to be suitable for the applications of nanocars [17,18,19], photoswitches [20], molecular heterojunctions [21], and catalysts [22]. Unusual molecular properties of multi-cage fullerene objects were theoretically predicted [23,24,25,26]. Recently, similar photophysical chemistry was also simulated in the modulation of photoswitchable dielectric properties to observe a large amplification of dielectric constants in a material combination form of multi-layered core-shell nanoparticles (NPs) [27,28,29]. The latter system was based on photoinduced intramolecular electronic charge-polarization of light-harvesting chromophoric nano[60]fullerenyl conjugates, such as 9,9-di(3,5,5-trimethylhexyl)-2-diphenylaminofluorenyl-methano[60]fullerene C60(>DPAF-C9) (1-C9, Figure 1). The polarization provided detectable dielectric property enhancement in a layered [60]fullerosome membrane structure on gold-shelled nanoparticles. The phenomena were based on the high electronegativity of C60> cage making it possible to rapidly shift an electron from light-harvesting DPAF (diphenylaminofluorene) donor moiety within the molecular structure to the C60> moiety. This resulted in the formation of the corresponding charge-separated (CS) state C60·(>DPAF-C9)+ as the source of polarized charges. In fact, ultrafast concurrent intramolecular energy and electron transfer kinetics within 1-C9 were substantiated previously by femtosecond transient absorption measurements (pump-probe) [30,31,32]. When these negative charges were distributed, delocalized, and stabilized in the fullerosome membrane array at the shell layer on core-shell NPs, it resulted in a CS state with a lifetime prolonged enough for the detection of dielectric characteristics. The process involved interlayer photoinduced plasmonic energy transfer from the Au shell layer to the outer shell layer of C60(>DPAF-C9) in addition to the fact that 1-C9 itself is also photoresponsive and excitable under light irradiation.
One crucial parameter to consider is the method of molecular packing within the fullerosome shell layer. In this regard, strong tendency of light-harvesting chromophores to aggregate among π-conjugated planar aromatic moieties can cause either concentration-dependent self-quenching effects of excited states or luminescence in the solid phase, including fullerosome. The ππ stacking may result in significant reduction of many photophysical properties. This type of packing aggregation can be partially minimized by the use of highly bulky and geometrically hindered π-conjugated chromophores in a structural design to restrict or distort intramolecular rotation bonding units with steric hindrance [33]. In the case of DPAF-C9, we recently developed and synthesized highly restricted 3D conformers based on inter-connected three DPAF-C9 chromophore units giving a structure of tris(DPAF-C9) (2-C9) [34] to prevent and minimize the tendency of planar DPAF units to undergo aromatic–aromatic stacking, overlapping, and aggregation via intermolecular hydrophobic–hydrophobic interactions in solid thin-films. This structural modification led the enhancement of photophysical properties, including the intensity of photoluminescence (PL) and electroluminescence (EL) emissions [35].
Accordingly, we extended the similar structural strategy to design new 3D conformers tris[(DPAF-C2M)-C60(>DPAF-C9)] (4-C2M–9), as shown in Figure 1, for the study. The stereochemical modification was based on the construction of 3D geometrically branched tris[(DPAF-C2M) (2-C2M) chromophore having a shared central benzene unit among three 2-diphenylaminofluorenes. Highly steric hindrance at the corresponding 1,3,5-phenylfluorenylaminobenzene moiety forced three fluorene ring moieties to twist either upward or downward away from the central benzene plane with a large torsional angle on the nitrogen atom. This resulted in the formation of three to four possible stereoisomeric configurations, such as cis-cup-2-C2M, trans-chair-2-C2M, and cis/or trans-propeller-2-C2M, as shown in Figure 1. All of these conformer forms were proposed to be capable of fully eliminating the tendency of 4-C2M–9 in inducing ππ aromatic–aromatic type stacking packing and to allow all C60 > cages to interact with each other via strong hydrophobic–hydrophobic interaction forces between (C60>)−(C60>) fullerenyl cages, forming the nano-layer array of fullerosome membrane.

2. Results and Discussion

Rapidly responsive nanophotonic physical properties of 3D conformers tris[(DPAF-C2M)-C60(>DPAF-C9)] (4-C2M−9) are achievable by specifically associating a donor–acceptor type chemical structure to a photomechanism having the ability to create a largely enhanced intramolecular energy and electron transfer efficiency. This mechanism occurs between C60> acceptor and DPAF-C2M/and DPAF-C9 donor moieties bonded on fullerenes. In our functional group design of 4-C2M−9, a methanoketo bridging unit was used to trigger the keto-enol isomerization tautomerism that is capable of inducing π-periconjugation between C60> and DFAP to provide a partial conjugation pathway for enhancing the π-electron mobility around the conjugated system of molecular nanostructures [36,37]. In addition, the new molecular design of stereoisomeric tris(fluorenylphenylamino)benzene [tris(DPAF-C9)] analogous was proven to act as a fluorophore showing high intensity of photoluminescence and electroluminescence emission efficiency [34,35]. This revealed the successful utilization of stereochemistry to allow hindered and branched 3′,5′,5′-trimethylhexyl (C9) arms to maintain the space-separation of three planar DPAF moieties intramolecularly within the nanostructure. It also behaved similarly in intermolecular packing that improved light-harvesting efficiency.
Based on the molecular formation energy based density functional theory (DFT) calculations of three plausible stereoisomers of tris(DPAF-C9) (2-C9) via B3LYP/6-31G* level of theory using SPARTAN08 [34,35], the results revealed high stability of the cis-cup-form with other forms in stability order of cup > chair > propeller, as shown in Figure 1. This agreed well for the alkyl n-C6, n-C7, and n-C8 substituents owing to the influence by strong dispersion interactions within the alkyl chains. In the case of the methyl and the ethyl substituents, trans-chair-form may have been more stable than cis-cup-form. Accordingly, three C4-analogous substituents of tris(DPAF-C2M) (2-C2M) facing toward each other in the 3D molecular space above the central benzene ring should have brought in the minimum alkyl–alkyl interaction forces required to keep a slight favor of the cis-cup-form over either trans-chair- or cis/trans-propeller form.
Synthetically, the precursor molecule 2-bromo-9,9-bis(methoxyethyl)fluorene (BrF-C2M) was prepared by alkylation of 2-bromofluorene with mesylated methoxyethanol reagent using potassium t-butoxide as a base in a high yield of 90% (Figure 1). The key 3D-conformeric intermediate 2-C2M was synthesized by the reaction of BrF-C2M with 1,3,5-tris(phenylamino)benzene (TPAB) in the presence of sodium t-butoxide, a catalytic amount of tris(dibenzylideneacetone)dipalladium(0) [Pd2(dba)3(0)], and rac-2,2′-bis(diphenylphosphino)- 1,1′-binaphthyl (rac-BINAP, 0.75 mol%) in anhydrous toluene at reflexing temperature for a period of 72 h to yield 82% of the product as a light yellow solid after chromatographic purification [SiO2, hexane–ethylacetate (1:1, v/v) as the eluent]. Subsequent attachment of three C60(>DPAF-C9) (1-C9) on 2-C2M should have led to the nanostructure of tris[(DPAF-C2M)-C60(>DPAF-C9)] (4-C2M–9) having all 1-C9 moieties extended outward from the central 1,3,5-triaminobenzene core. Prior to the attachment of three 1-C9 on 2-C2M, it was functionalized by the Friedel–Crafts acylation at C7 position of 2-diphenylaminofluorene moiety [36] with excessive α-bromoacetyl bromide (11.4 equiv.) in the presence of aluminum chloride (17 equiv.) in 1,2-dichloroethane at 0–10 °C to ambient temperature overnight to afford the corresponding α-bromoacetylfluorene derivative in this intermediate step of reactions. It resulted in viscous yellow semi-solids in 48% yield of tris(BrDPAF-C2M) (3-C2M). It was purified by either column or thin-layer chromatography (TLC) [silica gel, hexane–EtOAc (1:1, v/v) as eluent, Rf = 0.5 on TLC]. The final step of synthesis for the preparation of 3D-conformers 4-C2M–9 was performed by the reaction of 3-C2M with 1-C9 (5.0 equiv.) in the presence of 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) in toluene at room temperature for 8.0 h. During the first Bingel reaction period of 2.0 h, mono- and bis-adducts could be observed and detected by the TLC technique, showing two brown fraction bands (Rf = 0.2 and 0.3 on TLC plate, corresponding to mono- and bis-adducts, respectively). Subsequently, three brown fraction bands (Rf = 0.2, 0.3, and 0.4) appeared after 4 h of reaction, indicating sequential additions of C60(>DPAF-C9) on the starting substrate tris(BrDPAF-C2M) (3-C2M) with the Rf band at 0.2 gradually disappeared. At the end of the reaction, only two bands (Rf = 0.3 and 0.4) remained, with the latter being identified as a major product band of 4-C2M–9. After purification of this fraction by column chromatography (silica gel) using toluene–ethyl acetate (7:3) as the eluent, the brown solids of tris[(DPAF-C2M)-C60(>DPAF-C9)] were obtained in 79% yield.
The compound 4-C2M–9 exhibited good solubility in common organic solvents owing to its possession of three DPAF-C9 (with a total of six branched C9-alkyl groups) and three DPAF-C2M (with a total of six methoxyethyl groups) moieties (Figure 1), making three C60> cages become encapsulated in the center of the 3D molecular configuration. The 3D configuration of cis-cup form resulted in these alkyl groups being the main structural moieties interacting with the solvent. Accordingly, the compound had solubility (20 mg/mL in CHCl3 or CH2Cl2) over 10 times higher than that of C60 itself in toluene (1.4 mg/mL).

2.1. Spectroscopic Characterization of Synthetic 3D Configurated Fullerenyl Nanomaterials

All chemical conversions of intermediate chemicals to the corresponding products at each step of the reactions were characterized by various spectroscopic techniques. The functional attachment of three α-bromoaceto groups (3.0 equiv.) to tris(DPAF-C2M) given the product of tris(BrDPAF-C2M) was verified by both infrared (FT-IR) and 1H-NMR spectra. The former showed a new strong carbonyl (–C=O) stretching absorption band centered at 1673 cm−1, indicating each of the three carbonyl groups being bonded on a phenyl moiety, such as that of 2-C2M. This absorption wavelength was in clear contrast to the strong band absorption at 1725 cm−1 normally detectable for an alkyl carbonyl group. In the case of 1H-NMR spectrum, tris(BrDPAF-C2M) displayed characteristic new peak signals of three methylene protons (Hα) next to the carbonyl group of the α-bromoaceto moiety at δ 4.49 (Figure 2Ab) as compared with that of 2-C2M (Figure 2Aa). Subsequent attachment of three C60(>DPAF-C9) moieties to each of the three DPAF moieties of 3-C2M with a cylopropylaceto bridging unit on each C60> cage of three 1-C9 (applied as a reagent) showed evidence of changing solubility characteristics of the product 4-C2M–9 matching with those of 1-C9. Its FTIR spectrum displayed a slight shift of cyclopropyl keto group absorption band to υmax 1679 cm−1 (Figure 3d), which was assigned to the carbonyl (C=O) stretching band. It was also accompanied by an olefinic (C=C) absorption band centered at υmax 1591 cm−1. Both C=O and C=C bands were correlated to those of C60(>DPAF-C9) (Figure 3b) and tris[C60(>DPAF-C9)] (Figure 3c), showing a nearly identical absorption wavenumber. Most importantly, we were able to detect two typical fullerenyl cage bands at υmax 574 (w) and 524 (s) cm−1 (Figure 3d). These two bands were corresponding characteristic absorptions used to provide evidence of (C60>)-related monoadducts and bisadducts with absorption wavenumbers and relative intensity ratios differentiable from those of C60 (Figure 3a) and 1-C9 substituents (Figure 3b). Accordingly, we applied this IR technique for the product structure verification during the chemical conversion from 3-C2M to 4-C2M−9. Upon conversion of C60 to its monoadducts, such as those of Figure 3b,c, the remaining cage structure of C60> exhibited the same two bands with a reduced peak intensity for the 574 cm−1 band. The intensity of this band was further reduced in the structure of <C60>-like bisadduct, such as 4-C2M−9 (Figure 3d). Furthermore, the latter band at 524 cm−1 still remaining strong was indicative of successful attachment of C60(>DPAF-C9) moieties on 3-C2M due to the possibility of having the second malonate bridging unit being attached at or near the equator region of the C60> cage. This would have led to the retention of a C60 half-cage that enabled absorption at 524 cm−1.
By the analysis of 1H-NMR spectrum, disappearance of peaks corresponding to the chemical shift of α-proton (Hα) on the α-bromoaceto group of 3-C2M at δ 4.49 (Figure 2Ab) along with the appearance of new peaks over δ 5.25–5.78 (Figure 2Ac) provided evidence of successful formation of a cyclopropanyl keto-bridging unit between a C60> cage and the fluorene moiety. By using the previously reported chemical shift values of the α-proton (Hα’) in C60(>DPAF-C9) (Figure 2Ad) [20] and the related α-proton (Hα″) in tris(DPAF-C9) (2-C9, Figure 2Ae) [23] as the reference, we assigned these proton peaks to a combination of Hα and Hα’. A large downfield shift of the Hα chemical shift from those of 3-C2M at δ 4.49 to δ 5.25–5.78 for 4-C2M–9 provided clear evidence of three C60(>DPAF-C9) moieties being attached on the corresponding α-bromoaceto bridging units of 3-C2M. The characteristics of the multipeaks for Hα and Hα’ revealed a less symmetrical environment among these six protons of 4-C2M–9 as the geometric shape of the nanostructure extended to a 3D configuration. It is worthwhile to mention that a large down-fielded chemical shift value of either Hα’ or Hα’’ away from the normal value of δ 2.1–2.5 for an alkyl aceto-α-proton was caused by the influence of strong [60]fullerenyl ring current in close vicinity. In addition, the alkyl proton regions over δ 2.93 (methoxy proton, 18H) and δ 2.75–2.64 (ethylenoxy proton, 12H) of Figure 2Ac (marked by beige) were correlated well to those of 2-C2M at δ 2.91 (18H) and δ 2.64 (12H) (Figure 2Aa), respectively, indicating good retention of central tris(DPAF-C2M) core region without any structural change of methoxy groups during the Friedel–Crafts acylation reaction. It also showed a new group of methyl proton peaks at δ 0.30–2.07 having an integration ratio value of 113.28, which represented 114 fluorenyl protons of C9-alkyl proton (19Hs for each of the two C9-alkyls of DPAF-C9) and was indicative of six C9-alkyl groups in the structure, consistent with the product structure. Additional 1H-NMR spectroscopic data analyses on proton integrations of all proton peaks to substantiate and count for the molar quantity ratio among fluorene, methoxyethyl, and C9 alkyl moieties to prove the molecular formulation of 4-C2M–9 are provided in supporting information.
Most importantly, characteristics of central benzene protons at the core region could be used for the analysis of the relatively geometric configuration of three fluorenyl rings with respect to each other. With a symmetrical structure of TPAB, three benzene protons should have displayed a singlet peak in its 1H-NMR spectrum. Upon attachment of a bulky fluorenyl moiety at each diphenylamino group, it induced high torsional stress and steric hindrance at the nitrogen atom that forced each 9,9′-di(methoxyethyl)fluorene moiety to twist or rotate either upward or downward from the central benzene plane. The action resulted in two main 3D conformers: cis-cup-2-C2M and trans-chair-2-C2M. The former with three C60(>DPAF-C9) moieties facing upward on the same side in the structure gave a singlet Ha peak (Figure 1). The latter with one facing downward and two C60(>DPAF-C9) moieties facing upward in the structure resulted in two proton peaks for trans-Ha’ (1H) and trans-Hb’ (2H) (Figure 2Ba). By analyzing Figure 2Aa of tris(DPAF-C2M), a sharp singlet proton peak at δ 6.53 was assigned to the chemical shift of central benzene proton cis-Ha. This peak was compared with that of the Ha proton peak of cis-cup-tris(DPAF-C9) (cis-cup-2-C9, Figure 2Bb) showing even better resolution of the peak profile, indicating a high purity of one 3D conformer fraction in a cis-cup-2-C2M form. Surprisingly, this conformer fraction was, in fact, the major product. Apparently, the hydrophobic–hydrophobic dispersion interaction forces derived from three methoxyethyl chains and heteroatoms were stronger than those among all C4-alkyl groups, which led to higher tendency in formation of the cis-cup-form. Accordingly, subsequent attachment of three C60(>DPAF-C9) moieties on cis-cup-2-C2M led to a similar formation of corresponding cis-cup-tris[(DPAF-C2M)-C60(>DPAF-C9)] (cis-cup-4-C2M–9), all having 1-C9 moieties facing upward from the central benzene core at the same side with respect to each other. Additional structural analyses and discussions are provided in supporting information.
In the case of the potential formation of regio-isomers of 4-C2M−9 at the <C60> moiety, since the monoadduct structure of C60(>DPAF-C9) was well-defined, with the assistance of X-ray single crystal structural analysis of C60(>DPAF-C2) [31,36], its attachment on tris(BrDPAF-C2M) (3-C2M) was believed to be governed by the bulkiness and the steric hindrance of both relatively large entities to result in only a limited number of region isomers on the C60 cage. This was proven by the 1H-NMR spectrum of 4-C2M−9 showing only several Hα and Hα’ proton peaks at roughly δ 5.2–5.8 (Figure 2Ac) instead of the broad band normally seen for the existence of a large number of region isomers. To our surprise, a peak at 524 cm−1 assigned the characteristic infrared absorption band of a half-C60 cage (as stated above) showed close resemblance to those of the monoadduct C60(>DPAF-C9) (Figure 3b) and tris[C60(>DPAF-C9)] (Figure 3c) at an identical wavenumber. This implied the structure of the major regio-isomeric products had both addend moieties located at the same half-sphere of a C60 cage that left the other half-sphere of C60 untouched.

2.2. Photophysical and Physical Properties of 3D Conformeric Fullerenyl Nanomaterials

Photophysical properties of the 3D conformer cis-cup-4-C2M–9 were compared with those of precursor intermediates using the UV-vis spectroscopic technique. They were governed by two photoresponsive moieties, including three electron (e)-accepting fullerene cages and six light-harvesting DPAF antenna units as electron (e)-donors. The use of the latter was to enhance the optical absorption capability at longer visible wavelengths. The absorption wavelength could be varied and modulated by the appropriate chemical modification of functional substituents on fluorenyl moiety to affect electron-pushing (donating) and pulling (accepting toward the molecular edge of C60> cage moiety) mobility across the molecular π-conjugation system. As shown in Figure 4Ae of cis-cup-4-C2M–9, optical absorption of C60> cage moieties appeared mainly at the broad band centered at 296 nm (1.82 × 105 L mol−1 cm−1), whereas the band centered at 411 nm (1.11 × 105 L mol−1 cm−1) was attributed to the absorption of DPAF moieties. Characteristics of the latter band were compared with those of tris(DPAF-C2M) (Figure 4Aa), tris(BrDPAF-C2M) (Figure 4Ab), cis-cup-tris[C60(>DPAF-C9)] (Figure 4Ac), and C60(>DPAF-C9) (1-C9, Figure 4Ad), showing a clear bathochromic shift of the 351 nm band of 2-C2M to 406 nm (7.91 × 104 L mol−1 cm−1) of 3-C2M, which matched roughly with the 404 nm band of 1-C9 and the 402 nm band cis-cup-tris[C60(>DPAF-C9)] for the peak assignment. This assignment was also consistent with the observation of roughly equal absorption extinction coefficient (ε) values for 2-C2M, 3-C2M, and tris[C60(>DPAF-C9)] with the same three DPAF moieties per molecule. Upon the attachment of three C60(>DPAF-C9) moieties, optical absorptions of [60]fullerene moieties of cis-cup-4-C2M–9 (Figure 4Ae) at 296 nm became dominant in the spectrum with a higher ε value. It was accompanied by a weak characteristic (forbidden) steady state absorption band of the C60> moiety appearing at 692 nm (the insert of Figure 4A) with a slightly higher extinction coefficient for the monoadduct 1-C9 than the bisadduct cis-cup-4-C2M–9, which was also consistent with the photophysical property discussion above and provided further confirmation of a conjugated fullerenyl nanostructure.
In addition, a roughly 2.1-fold higher ε value (1.11 × 105 L mol−1 cm−1) of the 411 nm peak in Figure 4Ae compared to that of 404 nm band of cis-cup-tris[C60(>DPAF-C9)] was consistent with a double number of DPAF arms per molecule for the former. Furthermore, very efficient intramolecular energy transfer from the excited singlet state of DPAF-C9 antenna to C60> was detected, which nearly eliminated the fluorescence of C60(>DPAF-C9) (λex: 406 nm, Figure 4Bb). In high contrast, without any C60> cage in the structure, the compound of tris(DPAF-C2M) showed a strong intensity of fluorescence emission (λex: 352 nm, Figure 4Ba) that clearly indicated the loss of photoexcited energy being associated with the influence of [60]fullerene. With an additional DPAF-C9 antenna in the structure of 4-C2M–9, it began to experience a slightly excessive fluorescence emission (λex: 410 nm, Figure 4Bc) after the majority of photoexcited DPAF-C2M energy underwent direct intramolecular energy transfer to the closely bonded [60]fullerene cage.
In investigating the plausibility of photoinduced intramolecular electron (e)-transfer capability within the nanostructure of the 3D conformer cis-cup-tris[(DPAF-C2M)-C60(>DPAF-C9)] (cis-cup-4-C2M–9), we first investigated the unit character of redox potentials among all structural components, including the bisadduct-based <C60> cage and the DPAF-C9 moieties for comparison using the cyclic voltammetric (CV) technique. Several CV measurements were performed on the sample of cis-cup-4-C2M–9 in a solution of CH2Cl2 containing (n-butyl)4N+-PF6 as the electrolyte and Pt as both the working and the counter electrodes and with Ag/AgCl as the reference electrode.
To deliver appropriate redox potential analyses and data interpretation, related CV characteristics of a C60-bisadduct of C60(>t-Bu-malonate)2 with a <C60> cage attached by two t-butylmalonate groups and the precursor compound C60(>DPAF-C9) were collected. They were performed under the same CV condition over cyclic oxidation and reduction voltages versus Ag/Ag+ from −2.0 to 2.0 V as those for cis-cup-4-C2M–9, as shown in Figure 5. As a result, it displayed one reversible oxidation (1Eox of 1.51 V) reduction (1Ered of 1.32 V) cyclic wave with the first half wave oxidation potential (1E1/2,ox) of 1.42 V for the DPAF moieties of cis-cup-4-C2M–9 at positive voltages (Figure 5Ab,Bb). In the negative voltage region, its CV diagram displayed three reversible reductions at −0.34 (1Ered), −0.82 (2Ered), and −1.29 V (3Ered) with the corresponding cyclic oxidation waves at −0.15 (1Eox), −0.44 (2Eox), and −0.96 (3Eox), respectively. These data corresponded to the first to the third half wave reduction potentials of −0.25 (1E1/2,red), −0.63 (2E1/2,red), and −1.12 V (3E1/2,red), respectively. By comparison of these values to those of C60(>t-Bu-malonate)2 (Figure 5Aa,Ba) for the fullerene cage moiety and those of C60(>DPAF-C9) (Figure 5Ac,Bc) for both DPAF and fullerene cage moieties, highly consistent and reproducible redox potential characteristics among structural components were found that also substantiated the structural derivatization of tris[(BrDPAF-C2M) (3-C2M) with triple C60(>DPAF-C9) to form cis-cup-4-C2M–9. Accordingly, the latter exhibited combined CV characteristics of <C60> and DPAF-C9. These CV characteristics were reproducible for four repeated redox cycles with the reductive C60> and the oxidative DPAF potential profiles showing only slight changes at the potential range of −2.0 to 2.0 V. This implied good stability of the material under CV conditions that led to possible reuse of 4-C2M–9.

2.3. Evidence of Intramolecular Energy- and Electron-Transfer Events within cis-cup-4-C2M–9 by Detection of Corresponding Reactive Oxygen Species (ROS)

There is an appropriate approach to substantiate intramolecular energy and electron transfer events within the nanomolecular structure of cis-cup-4-C2M–9 by directly detecting the photoinduced production of reactive oxygen species (ROS). In general, the most common ROS includes singlet oxygen (1O2) produced by the Type-II photomechanism via the intermolecular transfer of triplet energy to molecular oxygen (O2) and superoxide radical (O2·) generated by the intermolecular transfer of electron (e) to O2. For the former case, upon photoexcitation at the C60> cage moiety of cis-cup-4-C2M–9, the singlet excited state of bis-methanofullerenyl 1(<C60>)* may undergo facile intersystem crossing in a nearly quantitative efficiency to its triplet excited state 3(<C60>)* that can be accounted for by the efficient production of 1O2 via Type-II triplet energy transfer processes. Alternatively, if the photoexcitation process is aimed at either the DPAF-C2M or the DPAF-C9 moiety of cis-cup-4-C2M–9, the resulting corresponding singlet excited states of either 1(DPAF-C2M)* or 1(DPAF-C9)* may undergo both pathways of (1) intramolecular energy transfer from either 1(DPAF-C2M)* or 1(DPAF-C9)* to the <C60> moiety to produce neutral DPAF-C2M or DPAF-C9 and 1(<C60>)*, respectively; (2) intramolecular electron (e)-transfer from either 1(DPAF-C2M)* or 1(DPAF-C9)* to the <C60> moiety to produce cationic (DPAF-C2M)+· or (DPAF-C9)+· and (<C60>)·, respectively. Both events of (1) and (2) can occur concurrently. Subsequent intermolecular e-transfer from (<C60>) to O2 produces the corresponding neutral <C60> and O2 following the Type-I photomechanism.
Accordingly, by the direct detection of ROS on either 1O2 and/or O2· upon irradiation on cis-cup-4-C2M–9 at either <C60> or DPAF-C2M/DPAF-C9 moiety, we were able to provide the evidence of intramolecular energy and electron transfer processes happening within this 3D-conformer. We selected two reliable fluorescent (FL) probes for the detection of either 1O2 or O2·separately in the solution of cis-cup-4-C2M–9 with high selectivity and specificity as a crucial measure. To detect the former ROS 1O2, a synthetic highly fluorescent compound α,α’-(anthracene-9,10-diyl)-bis(methylmalonic acid) (ABMA) was used as the probe in the experiment. Its UV-vis absorption and fluorescence emission spectra are given in Figure 6a,b, respectively. In the probe reaction, chemical trapping of 1O2 by highly fluorescent ABMA resulted in the formation of non-fluorescent 9,10-endoperoxide product ABMA-O2 (Figure 7A). This chemical conversion allowed us to follow the intensity loss of fluoresce emission upon photoexcitation. The loss could be associated with the proportional quantity of 1O2 produced. The correlation was valid owing to a higher reaction kinetic rate of the trapping process in solution than the internal decay of 1O2 in the same solvent system of a DMF–CHCl3 (1:9, v/v) mixture. Experimentally, the quantity of 1O2 generated was monitored and counted by the relative intensity decrease of fluorescence emission (λem) of ABMA at 428 nm under excitation wavelength (λex) of 380 nm. At this excitation wavelength, it matched partially with the optical absorption band of DPAF moieties of cis-cup-4-C2M–9 that led to a slight fluorescence emission (Figure 6c) after the intramolecular energy and the e-transfer processes. It gave a slightly higher count in the overall FL intensity during the experiment (Figure 6d). In a typical probe reaction, a master solution of ABMA in DMF was diluted by CHCl3 prior to the addition of cis-cup-4-C2M–9. It was followed by periodical illumination using a light emitting diode (LED) lamp of white light (a power output of >2.0 W) operated at two major emission peak maxima (λmax) centered at 451 and 530 nm. The former light emission spectrum exhibited a sufficient bandwidth covering the 410–470-nm region for photoexcitation of DPAF moieties with optical absorption bands covering 380–500-nm (Figure 4Ae). As a result, we were able to detect rapid production of 1O2 by cis-cup-4-C2M–9 upon irradiation in a decreasing curve profile over a period of more than 120 min (Figure 7Ab).
The FL probe experiments were calibrated by a blank control run using the same probe concentration of ABMA alone and an illumination time scale in the absence of cis-cup-4-C2M–9 (Figure 7Aa). Apparently, we observed slight photodegradation of ABMA itself. This may have implied the existence of a photoinduced triplet state of ABMA in a low quantity due to exposure to short wavelength regions of the light emission bandwidth covering over ~380 nm of ABMA absorption bands.
In the case of detecting superoxide radical as the second ROS, a synthetic O2·-reactive fluorescent probe precursor molecule, non-fluorescent potassium bis(2,4-dinitrobenzenesulfonyl)-2′,4′,5′,7′-tetrafluorofluorescein-10′ (or 11′)-carboxylate (DNBs-TFFC), was applied for the experiment. Its molecular structure was synthetically modified from that reported previously [38], showing good reaction selectivity with a high O2·/1O2 sensitivity ratio. Since DNBs-TFFC itself is photodegradable, a dialysis film with the molecular weight cut-off (MWCO) of 100−500 Daltons was applied to hold the solution of cis-cup-4-C2M–9 in toluene–DMSO (9:1, v/v). The sack bag was separated from the solution of the probe DNBs-TFFC. The latter was kept in a cuvette with stirring during the fluorescence emission measurement. Only the solution of cis-cup-4-C2M–9 in the dialysis membrane sack was subjected to the white LED light exposure. Any superoxide radical produced was allowed to rapidly diffuse into the probe solution through the dialysis membrane and initiate the desulfonylation of DNBs-TFFC. The O2·-trapping reaction led to the elimination of two dinitrobenzenesulfonyl moieties and yielded the corresponding bisphenol intermediate, as shown in Figure 7B. Rearrangement of the bisphenol intermediate to the ring-opening of lactone afforded highly fluorescent potassium 2′,4′,5′,7′-tetrafluorofluorescein-10′ (or 11′)-carboxylate regiosisomers (TFFC). The latter compound gave the fluorescence emission at 530 nm (λem) with the excitation at 484 nm (λex). As the probe DNBs-TFFC was not a fluorescent compound, detected emission photon counts were fully associated with the quantity of TFFC produced. Measured total emission intensity counts were then correlated to the relative quantity of O2· generated. As shown in Figure 7Bb, nearly linear progressive increase of fluorescence intensity counts over the full irradiation period was observed that revealed the constant production of O2· from the photoexcited cis-cup-4-C2M–9. As discussed above, continuous irradiation on six DPAF moieties of cis-cup-4-C2M–9 by white LED light (2.0 W) stimulated photoexcitation from the ground to the singlet excited state. Subsequent intramolecular e-transfer from 1(DPAF-Cn)* to <C60> moieties resulted in the formation of anionic [60]fullerenyl bisadduct radical (<C60>) intermediate. In the presence of O2, it was followed by further e-transfer from (<C60>) intermediate to O2 to produce O2 in a sequential multiple-step Type-I photomechanism. These results clearly provided the evidence of photoinduced intramolecular e-transfer mechanism within the 3D-conformer cis-cup-4-C2M–9.
Furthermore, one of the main crucial criterion for the use of these types of C60-(light-harvesting antenna)n conjugates, such as 4-C2M–9, as the nano-photosensitizers for antibacterial inactivation (aPDI) is their high photostability. Unlike the conventional organic chromophore-based photosensitizers suffering rapid photodegradation, C60-(light-harvesting antenna)n based nano-drugs were found to be capable of a single dose with multiple aPDI/PDT (photodynamic therapy) treatments [39,40,41,42].

3. Experimental Section

3.1. Chemicals and Reagents

Reagents of sodium t-butoxide, α-bromoacetyl bromide, aluminum chloride (AlCl3), 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU), rac-2,2′-bis(diphenylphosphino)-1,1′-binaphthyl (BINAP), and tris(dibenzylideneacetone)dipalladium(0) [Pd2(dba)3(0)] were purchased from Aldrich Chemicals, Natick, Massachusetts, USA and used without further purification. Fullerene materials with a purity of 99% were purchased from Suzhou Dade Carbon Nanotechnology Co., Ltd. Suzhou, Jiangsu, China. Anhydrous grade solvent of tetrahydrofuran (THF) was used and further dried via refluxing over sodium and benzophenone overnight and distilled under reduced pressure (10−1 mmHg). The precursor compounds including 1,3,5-tris(N-phenylamino)benzene (TPAB) and 2-bromo-9,9-di(methoxyethyl)fluorene (BrF-C2M) were synthesized according to our previous procedures [34].

3.2. Instruments for Spectroscopic Measurements

1H-NMR spectra were recorded on either Bruker and Spectrospin Avance 500 or Bruker AC-300 spectrometer. UV-vis spectra were recorded on a PerkinElmer Lambda 750 UV spectrometer. Fluorometric traces were collected using a PTI QuantaMasterTM40 Fluorescence Spectrofluorometer. The light source used in this experiment included a collimated white LED light with an output power of 2.0 W (Prizmatix, Southfield, MI, USA). Infrared spectra were recorded as KBr pellets on a Thermo Nicolet AVATAR 370 FTIR spectrometer. Cyclic voltammetry (CV) was record on EG&G Princeton Applied Research 263A Potentiostat/Galvanostat using Pt metal as the working electrode, Ag/AgCl as the reference electrode, and Pt wire as the counter electrode at a scan rate of 10 mV/s. The solution for CV measurements was prepared in a concentration of 1.0–5.0 × 10−3 M in appropriate solvents containing the electrolyte Bu4N+-PF6 (0.1 M).

3.3. Synthesis of N1,N3,N5-Tris(9,9-di(methoxyethyl)fluoren-2-yl)-1″,3″,5″-tris(phenylamino)-benzene as Tris(DPAF-C2M) (2-C2M)

Synthetic procedure for the preparation of tris(DPAF-C2M) was slightly modified from those methods reported recently [34]. In general, a mixture of BrF-C2M (7.33 g, 20.3 mmol, excess), TPAB (1.16 g, 3.3 mmol), and sodium t-butoxide (1.94 g, 20.3 mmol) was dissolved in anhydrous toluene (75 mL) and stirred for 1 h to give a homogeneous solution. The catalyst Pd2(dba)3(0) (0.023 g, 0.25 mol%) and rac-BINAP (0.046 g, 0.75 mol%) were added to the solution, followed by heating to refluxing temperature under nitrogen for a period of 72 h. After cooling the resulting mixture to room temperature, it was washed with water three times by extraction, the organic layer was separated, and it was dried over sodium sulfate. After solvent evaporation, a small quantity of crude paste was tested on the TLC plate to show the major product at Rf = 0.6 using hexane–ethylacetate (1:1, v/v) as the eluent. This product spot had a dense yellow-brown color on the top accompanied by a light visible tail. The tail portion was assumed to be the product in the trans-chair form. This tail portion was subsequently separated from the main top portion of the cis-cup form via column chromatography, followed by the TLC plate purification using silica gel as the stationary phase and hexane-ethylacetate (1:1, v/v) as the eluent to afford the product of cis-cup-tris(DPAF-C2M) (2-C2M) as light yellow solids in 82% yield (3.23 g). Verification of the cis-cup form was based on the detection of a singlet proton peak at δ6.53 (Ha) corresponding to three central core 1,3,5-benzene protons, whereas the trans-chair form gave two proton groups at a integration ratio of 2:1. Spectroscopic data: FT-IR (KBr) υmax 3062 (w, aromatic C-H stretching), 3031 (w), 3016 (w), 2969 (w, aliphatic C-H stretching), 2922 (m), 2875 (m), 2814 (w), 1571 (s, C=C), 1491 (s, anti-symmetric deformations of CH3 groups and scissor vibrations of CH2 groups), 1448 (s), 1428 (m), 1381 (w, symmetric deformations of CH3 groups), 1292 (m, asymmetric stretching vibrations of C-N-C), 1248 (m, asymmetric stretching vibrations of C-N-C), 1212 (w), 1176 (w), 1155 (w), 1110 (s, stretching vibrations of C-O-C), 1039 (w), 948 (w), 877 (w), 829 (w), 754 (m), 738 (s, out-of-plan deformation of C-H), 711 (m, out-of-plan deformation of C-H), 692 (s), and 511 (w) cm−1; UV-vis (CHCl3, 1.0 × 10−5 M) λmax (ε) 321 (7.24 × 104 L mol−1 cm−1) and 351 nm (7.66 × 104 L mol−1 cm−1); 1H NMR (500 MHz, CDCl3) δ 7.58 (s, 3H, br), 7.52 (d, 3H), 7.35–7.24 (m, 9H), 7.15–7.03 (m, 18H), 6.89 (t, 3H), 6.53 (s, 3H, central benzene protons Ha), 2.91 (s, 18H), 2.65 (m,12H), and 2.19 (m, 12H).

3.4. Synthesis of N1,N3,N5-Tris(7-α-bromoacetyl-9,9-di(methoxyethyl)fluoren-2-yl)-1″,3″,5″-tris-(phenylamino)benzene as Tris(BrDPAF-C2M) (3-C2M)

The compound cis-cup-tris(DPAF-C2M) (cis-cup-2-C2M, 0.53 g, 0.44 mmol) was added to a homogeneous suspension of AlCl3 (1.0 g, 7.5 mmol) in 1,2-dichloroethane (40 mL) at 0 °C with vigorous stirring. The reagent α-bromoacetyl bromide (1.0 g, 5.0 mmol) was added slowly over 10 min while maintaining the temperature between 0–10 °C. The mixture was then stirred overnight at room temperature. An excessive amount of AlCl3 remaining in the solution was quenched by slow addition of water (50 mL) while maintaining the temperature below 45 °C. After washing sequentially with dil. HCl (1.0 N, 50 mL) and water (50 mL × 2), the organic layer was separated and dried over sodium sulfate and then concentrated in vacuo to give the crude product as viscous yellow semi-solids. It was purified by column chromatography (silica gel) followed by thin-layer chromatography (TLC) using hexane–EtOAc (1:1, v/v) as eluent to afford cis-cup-tris(Br-DPAF-C2M) (cis-cup-3-C2M) (at Rf = 0.5 on TLC plate) in 48% yield (0.33 g). Spectroscopic data: FT-IR (KBr) υmax 3062 (w, aromatic C-H stretching), 3031 (w), 3016 (w), 2969 (w, aliphatic C-H stretching), 2922 (m), 2875 (m), 1673 (s, C=O), 1571 (s, C=C), 1491 (s, anti-symmetric deformations of CH3 groups and scissor vibrations of CH2 groups), 1467 (s), 1448 (s), 1428 (m), 1388 (w), 1348 (w), 1282 (s), 1251 (s), 1195 (m), 1176 (w), 1110 (s, stretching vibrations of C-O-C), 1035 (w), 948 (w), 879 (w), 823 (m), 755 (m), 740 (s, C-H out-of-plan deformation), 715 (m, C-H out-of-plan deformation), 694 (s), 617 (w) and 538 (w) cm−1; UV-vis (CHCl3, 1.0 × 10−5 M) λmax (ε) 310 (7.65 × 104 L mol−1 cm−1) and 406 nm (7.91 × 104 L mol−1 cm−1); 1H NMR (500 MHz, CDCl3) δ 7.98 (m, 3H), 7.58 (m, 3H), 7.34–7.20 (m, 9H), 7.11–6.98 (m, 18H), 6.96 (m, 3H), 6.56 (m, 3H, aromatic protons of central phenyl ring), 4.49 (m, 6H, α-proton next on C61), 2.91 (s, 18H. primary C2M alkyl protons next to O-atom), 2.81–2.50 (centered at δ 2.66, m, 12H, secondary C2M alkyl protons next to O-atom), 2.48–2.07 (centered at δ 2.16, m, 12H, C2M alkyl protons next to the fluorene ring).

3.5. Synthesis of N1,N3,N5-Tris(7-(1,2-dihydro-1,2-methanofullerene[60]-61-carbonyl)-9,9-di(methoxyethyl)fluoren-2-yl)-1″,3″,5″-tris(phenylamino)benzene) as Tris[(DPAF-C2M)-C60(>DPAF-C9)] (4-C2M–9)

The reagent 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU, 0.21 g, 1.38 mmol) was added slowly to a homogeneous mixture of C60>(DPAF-C9) (1.1 g, 0.97 mmol) and cis-cup-tris(BrDPAF-C2M) (cis-cup-3-C2M, 0.31 g, 0.20 mmol) in anhydrous toluene (700 mL). During the Bingel reaction for the first 2.0 h, the products of mono- and bis-adducts became visible by the TLC technique, showing two brown bands at Rf = ~0.2 and ~0.3, respectively, using a mixture of toluene–EtOAc (7:3, v/v) as eluent. After a longer reaction period of 4 h, three brown bands close to each other were observed at Rf = ~0.2, ~0.3, and ~0.4 with the former band progressively becoming faint. At the end of the reaction (8.0 h), only two latter bands remained at Rf = ~0.3 and ~0.4 with the latter as the major chromatographic fraction. At this stage, the reaction mixture was concentrated to a 10% volume and then precipitated from methanol (100 mL) to afford the crude product mixture, which was isolated by centrifugation. Further purification by column chromatography (silica gel) using toluene to a solvent mixture of toluene–EtOAc (7:3, v/v) as the eluent with sequential increments of increasing solvent polarity afforded cis-cup-tris[(DPAF-C2M)-C60(>DPAF-C9)]. It was further purified by TLC with isolation of only the narrow dense color fraction band to give brown solids of cis-cup-4-C2M–9 in 79% yield (0.74 g) (at Rf = ~0.4 on TLC). Spectroscopic data: FT-IR (KBr) υmax 3062 (w, aromatic C-H stretching), 3031 (w), 3016 (w), 2952 (w, aliphatic C-H stretching), 2925 (m), 2852 (w), 1679 (s, C=O), 1591 (s, C=C), 1490 (s, anti-symmetric deformations of CH3 groups and scissor vibrations of CH2 groups), 1465 (s), 1419 (m), 1346 (w), 1315 (m), 1274 (s), 1238 (m), 1211 (s), 1170 (m), 1110 (s, stretching vibrations of C-O-C), 1072 (w), 1029 (w), 950 (w), 879 (w), 815 (m), 746 (s), 694 (s), 574 (w) and 524 (s, <C60>) cm−1; UV-vis (CHCl3, 1.0 × 10−5 M) λmax (ε) 296 nm (1.82 × 105 L mol−1 cm−1) and 411 nm (1.11 × 105 L mol−1 cm−1); 1H NMR (500 MHz, CDCl3) δ 8.65–8.12 (m, 12H, fluorenyl protons next to the keto group), 7.91–7.48 (m, 12H, fluorenyl protons), 7.31−7.06 [m, 57H, 45 aminophenyl protons and 12 fluorenyl protons (2H for each fluorene ring) next to N-atom], 6.57 (m, 3H, aromatic protons of the central phenyl ring), 5.78–5.25 (m, 6H, α-proton next on C61), 3.01–2.81 (centered at δ 2.93) (m, 18H, primary C2M alkyl protons next to O-atom), 2.81–2.50 (centered at δ 2.65, m, 12H, secondary C2M alkyl protons next to O-atom), 2.48–2.07 (centered at δ 2.17, m, 12H, C2M alkyl protons next to the fluorine ring), 2.07−1.18 (m, 12H, C9 alkyl protons next to the fluorene ring), 1.18–0.94 (centered at δ 1.14) (m, 6H, tertiary C9 alkyl protons), 0.94–0.30 (centered at δ 0.70) (m, 96H, primary and secondary C9 alkyl protons).

3.6. ROS Measurements Using singlet oxygen (1O2)-Sensitive Fluorescent Probe

The compound α,α’-(anthracene-9,10-diyl)bis(methylmalonic acid) (ABMA) was used as a fluorescent probe for singlet oxygen (1O2) trapping. The quantity of 1O2 generated was monitored and counted by the relative intensity decrease of fluorescence emission of ABMA at 428 nm under excitation wavelengths of 380 nm (λex). A typical probe solution was prepared by diluting a master solution of ABMA (1.0 × 10−5 M in DMF, 0.4 mL) with an amount 9-fold in volume of CHCl3 (3.2 mL) in a cuvette (10 × 10 × 45 mm). The solution was added by a pre-defined volume of tris[(DPAF-C2M)-C60(>DPAF-C9)] in CHCl3 (1.0 × 10−5 M, 0.4 mL), followed by periodic illumination using an ultrahigh power white light LED lamp (Prizmatix, operated at the emission peak maxima centered at 451 and 530 nm with the collimated optical power output of >2.0 W in a diameter of 5.2 cm). Progressive fluorescent spectra were taken on the PTI QuantaMasterTM 40 Fluorescence Spectrofluorometer.

3.7. ROS Measurements Using Superoxide Radical (O2·)-Sensitive Fluorescent Probe

A superoxide radical (O2∙)-reactive fluorescent probe, non-fluorescent potassium bis(2,4-dinitrobenzenesulfonyl)-2′,4′,5′,7′-tetrafluorofluorescein-10′ (or 11′)-carboxylate regioisomers (DNBs-TFFC), was used for the experiment. A typical probe solution [10−6 M in toluene–DMSO (9:1)] was prepared by diluting a stock solution of DNBs-TFFC in DMSO (10−5 M, 1.0 mL) by 10 times with toluene (9.0 mL). A dialysis film (CE) with the molecular weight cut-off (MWCO) of 100–500 Da was used to separate the solution of tris[(DPAF-C2M)-C60(>DPAF-C9)] [10−6 M in toluene–DMSO (9:1)] from the probe solution kept in a cuvette with stirring during the fluorescent measurement. Only the solution of 4-C2M in the membrane sack was subjected to the LED light exposure at the excitation wavelength of 400–700 nm (white light). The quantity of O2· generated was counted in association with its reaction with DNBs-TFFC that resulted in the product of highly fluorescent potassium 2′,4′,5′,7′-tetrafluorofluorescein-10′ (or 11′)-carboxylate regioisomers (TFFC) with fluorescence emission at 530 nm (λem) upon excitation at 484 nm (λex). The detected intensity increase of fluorescence emission was then correlated to the relative quantity of O2 produced.

4. Conclusions

Previous studies on detected photoswitchable dielectric amplification phenomena by the simulated ferroelectric-like capacitor design using the construction and the fabrication of a fullerosome shell layer on core-shell γ-FeOx@AuNPs were based on the photoinduced intramolecular charge-polarization of (C60> acceptor)-(DPAF-Cn donor) conjugates [16–18]. The corresponding formation of dielectric ion-radical components (C60>) and DPAF+·-Cn within a fullerosome array layer was the basis of observed dielectric properties. Accordingly, several such conjugates were developed by the extension from the initial C60(>DPAF-C9) to demonstrate the correlation of the structural relationship and the chemical modifications to the enhanced dielectric properties, as stated above. Two interesting modifications both involved 3D-conformeric C60(>DPAF-C9) derivatives by fusing three phenyl rings of three diphenylamino groups together to form a shared central benzene moiety as the base of 3D configuration design. Specifically, the successful synthesis of cis-cup-tris[(DPAF-C2M)-C60(>DPAF-C9)] stereoisomer may be beneficial for use as positive (DPAF-Cn)+ and negative charge (<C60>) carriers in enhancing photoinduced dielectric characteristics [43]. It can also be applied as the precursor building block in the synthesis of several C60- and C70-based ultrafast photoresponsive nonlinear two-photon absorptive nanomaterials. Accordingly, we demonstrated efficient intramolecular energy and electron transfer capabilities of 3D conformer cis-cup-4-C2M–9 using photophysical measurements and its effective production of singlet oxygen (via the energy transfer mechanism) and superoxide radicals (via the electron transfer mechanism). They can be applied as nano-photosensitizers [39,40,41,42] and nonlinear photonic agents [30,31,32,44].

Supplementary Materials

Supplementary Materials can be accessed online.

Author Contributions

All authors contributed significant effort on this work. H.Y. and M.W. carried out the main synthetic works, spectroscopic characterization, data analysis, and physical measurements; L.-S.T., and L.Y.C. participated in the discussion and experimental studies and contributed to a part of manuscript writing; All authors read and approved the final manuscript.

Funding

This research was funded by Air Force Office of Scientific Research (AFOSR) under the grant number FA9550-14-1-0153.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. El-Khouly, M.E.; Ito, O. Intermolecular and supramolecular photoinduced electron transfer processes of fullerene–porphyrin/phthalocyanine systems. J. Photochem. Photobiol. C Photochem. Rev. 2004, 5, 79–104. [Google Scholar]
  2. Escudero, D. Revising intramolecular photoinduced electron transfer (PET) from first-principles. Acc. Chem. Res. 2016, 49, 1816–1824. [Google Scholar] [CrossRef] [PubMed]
  3. Ito, O.; D’Souza, F. Recent advances in photoinduced electron transfer processes of fullerene-based molecular assemblies and nanocomposites. Molecules 2012, 17, 5816–5835. [Google Scholar] [PubMed]
  4. Wu, W.; Zhao, J.; Sun, J.; Guo, S. Light-harvesting fullerene dyads as organic triplet photosensitizers for triplet–triplet annihilation upconversions. J. Org. Chem. 2012, 77, 5305–5312. [Google Scholar] [CrossRef] [PubMed]
  5. Ziessel, R.; Allen, B.D.; Rewinska, D.B.; Harriman, A. Selective triplet-state formation during charge recombination in a fullerene/bodipy molecular dyad (bodipy=borondipyrromethene). Chem. Eur. J. 2009, 15, 7382–7393. [Google Scholar] [CrossRef] [PubMed]
  6. Zhao, J.; Wu, W.; Sun, J.; Guo, S. Triplet photosensitizers: From molecular design to applications. Chem. Soc. Rev. 2013, 42, 5323–5351. [Google Scholar] [CrossRef]
  7. Chea, Y.; Yuan, X.; Cai, F.; Zhaoa, J.; Zhaoa, X.; Xub, H.; Liu, L. Bodipy−corrole dyad with truxene bridge: Photophysical properties and application in triplet−triplet annihilation upconversion. Dyes Pigments 2019, 171, 107756. [Google Scholar] [CrossRef]
  8. Kamkaew, A.; Lim, S.H.; Lee, H.B.; Kiew, L.V.; Chung, L.Y.; Burgess, K. BODIPY dyes in photodynamic therapy. Chem. Soc. Rev. 2013, 42, 77–88. [Google Scholar] [CrossRef]
  9. Natali, M.; Campagna, S.; Scandola, F. Photoinduced electron transfer across molecular bridges: Electron- and hole-transfer superexchange pathways. Chem. Soc. Rev. 2014, 43, 4005–4018. [Google Scholar] [CrossRef]
  10. Imahori, H.; Sakata, Y. Donor-linked fullerenes: Photoinduced electron transfer and its potential application. Adv. Mater. 1997, 9, 537–546. [Google Scholar]
  11. D’Souza, F.; Ito, O. Photosensitized electron transfer processes of nanocarbons applicable to solar cells. Chem. Soc. Rev. 2012, 41, 86–96. [Google Scholar] [CrossRef] [PubMed]
  12. Bottari, G.; Torre, G.; Guldi, D.M.; Torres, T. Covalent and noncovalent phthalocyanine-carbon nanostructure systems: Synthesis, photoinduced electron transfer, and application to molecular photovoltaics. Chem. Rev. 2010, 110, 6768–6816. [Google Scholar] [CrossRef] [PubMed]
  13. Daly, B.; Ling, J.; Prasanna de Silva, A. Current developments in fluorescent PET (photoinduced electron transfer) sensors and switches. Chem. Soc. Rev. 2015, 44, 4203–4211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Yin, R.; Wang, M.; Huang, Y.-Y.; Chiang, L.Y.; Hamblin, M.R. Photodynamic therapy with decacationic [60]fullerene monoadducts: Effect of a light absorbing e-donor antenna and micellar formulation. Nanomed. Nanotechnol. Biol. Med. 2014, 10, 795–808. [Google Scholar] [CrossRef] [PubMed]
  15. Yin, R.; Wang, M.; Huang, Y.-Y.; Landi, G.; Vecchio, D.; Chiang, L.Y.; Hamblin, M.R. Antimicrobial photodynamic inactivation with decacationic functionalized fullerenes: Oxygen independent photokilling in presence of azide and new mechanistic insights. Free Radic. Biol. Med. 2014, 79, 14–27. [Google Scholar] [CrossRef] [PubMed]
  16. Segura, J.L.; Martin, N. [60]Fullerene dimer. Chem. Soc. Rev. 2000, 29, 13–25. [Google Scholar] [CrossRef]
  17. Shirai, Y.; Osgood, A.J.; Zhao, Y.; Kelly, K.F.; Tour, J.M. Directional control in thermally driven single-molecule nanocars. Nano Lett. 2005, 5, 2330–2334. [Google Scholar] [CrossRef]
  18. Akimov, A.V.; Nemukhin, A.V.; Moskovsky, A.A.; Kolomeisky, A.B.; Tour, J.M. Molecular dynamics of surface-moving thermally driven nanocars. J. Chem. Theory Comput. 2008, 4, 652–656. [Google Scholar] [CrossRef]
  19. Sasaki, T.; Osgood, A.J.; Kiappes, J.L.; Kelly, K.F.; Tour, J.M. Synthesis of a porphyrin-fullerene pinwheel. Org. Lett. 2008, 10, 1377–1380. [Google Scholar] [CrossRef]
  20. Zhang, J.; Porfyrakis, K.; Morton, J.J.L.; Sambrook, M.R.; Harmer, J.; Xiao, L.; Ardavan, A.; Briggs, G.A.D.; Briggs, G. Photoisomerization of a fullerene dimer. J. Phys. Chem. 2008, C 112, 2802–2904. [Google Scholar] [CrossRef]
  21. Wang, J.L.; Duan, X.F.; Jiang, B.; Gan, L.B.; Pei, J.; He, C.; Li, Y.F. Nanosized rigid π-conjugated molecular heterojunctions with multi[6 0]fullerenes: Facile synthesis and photophysical properties. J. Org. Chem. 2006, 71, 4400–4410. [Google Scholar] [CrossRef] [PubMed]
  22. Lόpez-Andarias, J.; Bauza, A.; Sakai, N.; Frontera, A.; Matile, S. Remote control of anion-π catalysis on fullerene-centered catalytic triads. Angew. Chem. 2018, 130, 11049–11053. [Google Scholar] [CrossRef]
  23. Sabirov, D.S. Polarizability of C60 fullerene dimer and oligomers: The unexpected enhancement and its use for rational design of fullerene-based nanostructures with adjustable properties. RSC Adv. 2013, 3, 19430. [Google Scholar] [CrossRef]
  24. Pankratyev, E.Y.; Tukhbatullina, A.A.; Sabirov, D.S. Dipole polarizability, structure, and stability of [2+2]-linked fullerene nanostructures (C60)n (n ≤ 7). Phys. E Low-Dimens. Syst. Nanostruct. 2017, 86, 237–242. [Google Scholar] [CrossRef]
  25. Tukhbatullina, A.; Shepelevich, I.; Sabirov, D.S. Exaltation of polarizability as a common property of fullerene dimers with divers intercage bridges. Fuller. Nanotub. Carbon Nanostruct. 2018, 26, 661–666. [Google Scholar] [CrossRef]
  26. Swart, M.; van Duijnen, P.T. Rapid determination of polarizability exaltation in fullerene-based nanostructures. J. Mater. Chem. 2015, C3, 23–25. [Google Scholar] [CrossRef]
  27. Wang, M.; Su, C.; Yu, T.; Tan, L.-S.; Hu, B.; Urbas, A.; Chiang, L.Y. Novel photoswitchable dielectric properties on nanomaterials of electronic core-shell γ-FeOx@Au@fullerosomes for GHz frequency applications. Nanoscale 2016, 8, 6589–6599. [Google Scholar] [CrossRef]
  28. Wang, M.; Yu, T.; Tan, L.-S.; Urbas, A.; Chiang, L.Y. Tunability of rf-responses by plasmonic dielectric amplification using branched e-polarizable C60-adducts on magnetic nanoparticles. J. Phys. Chem. C 2016, 120, 17711–17721. [Google Scholar] [CrossRef]
  29. Wang, M.; Yu, T.; Tan, L.-S.; Urbas, A.; Chiang, L.Y. Enhancement of photoswitchable dielectric property by conducting electron donors on plasmonic core-shell gold-fluorenyl C60 nanoparticles. J. Phys. Chem. C 2018, 122, 12512–12523. [Google Scholar] [CrossRef]
  30. Padmawar, P.A.; Rogers, J.O.; He, G.S.; Chiang, L.Y.; Canteenwala, T.; Tan, L.-S. Large cross-section enhancement and intramolecular energy transfer upon multiphoton absorption of hindered diphenylaminofluorene-C60 dyads and triads. Chem. Mater. 2006, 18, 4065–4074. [Google Scholar] [CrossRef]
  31. Padmawar, P.A.; Canteenwala, T.; Tan, L.-S.; Chiang, L.Y. Synthesis and characterization of photoresponsive diphenylaminofluorene chromophore adducts of [60]fullerene. J. Mater. Chem. 2006, 16, 1366–1378. [Google Scholar] [CrossRef]
  32. Luo, H.; Fujitsuka, M.; Araki, Y.; Ito, O.; Padmawar, P.; Chiang, L.Y. Inter- and intramolecular photoinduced electron-transfer processes between C60 and diphenylaminofluorene in solutions. J. Phys. Chem. B 2003, 107, 9312–9318. [Google Scholar] [CrossRef]
  33. Hu, R.; Lager, E.; Aguilar-Aguilar, A.; Liu, J.; Lam, J.W.Y.; Sung, H.H.Y.; Williams, I.D.; Zhong, Y.; Wong, K.S.; Pena-Cabrera, E.; et al. Twisted intramolecular charge transfer and aggregation-induced emission of BODIPY derivatives. J. Phys. Chem. C 2009, 113, 15845–15853. [Google Scholar] [CrossRef]
  34. Kang, N.-G.; Kokubo, K.; Jeon, S.; Wang, M.; Lee, C.-L.; Canteenwala, T.; Tan, L.-S.; Chiang, L. Synthesis and photoluminescent properties of geometrically hindered cis-tris(diphenyl-aminofluorene) as precursors to light-emitting devices. Molecules 2015, 20, 4635–4654. [Google Scholar] [CrossRef] [PubMed]
  35. Lee, Y.-T.; Wang, M.; Kokubo, K.; Kang, N.-G.; Wolf, L.; Tan, L.-S.; Chen, C.-T.; Chiang, L. New 3D-stereoconfigurated cis-tris(fluorenylphenylamino)-benzene with large steric hindrance to minimize ππ stacking in thin-film devices. Dyes Pigments 2018, 149, 377–386. [Google Scholar] [CrossRef]
  36. Chiang, L.Y.; Padmawar, P.A.; Canteenwala, T.; Tan, L.-S.; He, G.S.; Kannan, R.; Vaia, R.; Lin, T.-C.; Zheng, Q.; Prasad, P.N. Synthesis of C60-diphenylaminofluorene dyad with large 2PA cross-sections and efficient intramolecular two-photon energy transfer. Chem. Commun. 2002, 1854–1855. [Google Scholar] [CrossRef]
  37. Jeon, S.; Wang, M.; Ji, W.; Tan, L.-S.; Cooper, T.; Chiang, L.Y. Broadband two-photon absorption characteristics of highly photostable fluorenyl-dicyanoethylenylated [60]fullerene dyads. Molecules 2016, 21, 647. [Google Scholar] [CrossRef]
  38. Maeda, H.; Yamamoto, K.; Nomura, Y.; Kohno, I.; Hafsi, L.; Ueda, N.; Yoshida, S.; Fukuda, M.; Fukuyasu, Y.; Yamauchi, Y.; et al. A design of fluorescent probes for superoxide based on a nonredox mechanism. J. Am. Chem. Soc. 2005, 127, 68–69. [Google Scholar] [CrossRef]
  39. Wang, M.; Huang, L.; Sharma, S.K.; Jeon, S.; Thota, S.; Sperandio, F.F.; Nayka, S.; Chang, J.; Hamblin, M.R.; Chiang, L.Y. Synthesis and photodynamic effect of new highly photostable decacationically armed [60]- and [70]fullerene decaiodide monoadducts to target pathogenic bacteria and cancer cells. J. Med. Chem. 2012, 55, 4274–4285. [Google Scholar] [CrossRef]
  40. Wang, M.; Maragani, S.; Huang, L.; Jeon, S.; Canteenwala, T.; Hamblin, M.R.; Chiang, L.Y. Synthesis of decacationic [60]fullerene decaiodides giving photoinduced production of superoxide radicals and effective PDT-mediation on antimicrobial photoinactvation. Eur. J. Med. Chem. 2013, 63, 170–184. [Google Scholar] [CrossRef]
  41. Sperandio, F.F.; Sharma, S.K.; Wang, M.; Jeon, S.; Huang, Y.-Y.; Dai, T.; Nayka, S.; de Sousa, S.C.O.M.; Chiang, L.Y.; Hamblin, M.R. Photoinduced electron-transfer mechanisms for radical-enhanced photodynamic therapy mediated by water-soluble decacationic C70 and C84O2 fullerene derivatives. Nanomed. Nanotech. Biol. Med. 2013, 9, 570–579. [Google Scholar] [CrossRef] [PubMed]
  42. Huang, L.; Wang, M.; Huang, Y.-Y.; El-Hussein, A.; Chiang, L.Y.; Hamblin, M.R. Progressive cationic functionalization of chlorin derivatives for antimicrobial photodynamic inactivation and related vancomycin conjugate. Photochem. Photobiol. Sci. 2018, 17, 638–651. [Google Scholar] [CrossRef] [PubMed]
  43. Yin, H.; Wang, M.; Yu, T.; Tan, L.-S.; Chiang, L.Y. 3D-conformer of tris[60]fullerenylated cis-tris(diphenylaminofluorene) as photoswitchable charge-polarizer on GHz-responsive trilayered core-shell dielectric nanoparticles. Molecules 2018, 23, 1873. [Google Scholar] [CrossRef] [PubMed]
  44. Jeon, S.; Haley, J.; Flikkema, J.; Nalla, V.; Wang, M.; Sfeir, M.; Tan, L.-S.; Cooper, T.; Ji, W.; Hamblin, M.R.; et al. Linear and nonlinear optical properties of light-harvesting hybrid [60]fullerene triads and tetraads with dual NIR two-photon absorption characteristics. J. Phys. Chem. C 2013, 117, 17186–17195. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Sample of tris[(DPAF-C2M)-C60(>DPAF-C9)] is available from the authors.
Figure 1. Synthetic methods for the preparation of 3D conformers of tris[(DPAF-C2M)-C60(>DPAF-C9)] (4-C2M–9) with reaction reagents given and three perspective 3D-configurations of the key stereoisomeric intermediate 2-C2M.
Figure 1. Synthetic methods for the preparation of 3D conformers of tris[(DPAF-C2M)-C60(>DPAF-C9)] (4-C2M–9) with reaction reagents given and three perspective 3D-configurations of the key stereoisomeric intermediate 2-C2M.
Molecules 24 03337 g001
Figure 2. 1H-NMR spectra (CDCl3) of (A) (a) tris(DPAF-C2M) (2-C2M), (b) tris(BrDPAF-C2M) (3-C2M), (c) tris[(DPAF-C2M)-C60(>DPAF-C9)] (4-C2M−9), (d) C60(>DPAF-C9) (1-C9), and (e) tris[C60(>DPAF-C9)]. (B) 1H-NMR spectra (CDCl3) of (a) trans-chair-tris(DPAF-C9) (trans-chair-2-C9) and (b) cis-cup-tris(DPAF-C9) (cis-cup-2-C9) for comparison.
Figure 2. 1H-NMR spectra (CDCl3) of (A) (a) tris(DPAF-C2M) (2-C2M), (b) tris(BrDPAF-C2M) (3-C2M), (c) tris[(DPAF-C2M)-C60(>DPAF-C9)] (4-C2M−9), (d) C60(>DPAF-C9) (1-C9), and (e) tris[C60(>DPAF-C9)]. (B) 1H-NMR spectra (CDCl3) of (a) trans-chair-tris(DPAF-C9) (trans-chair-2-C9) and (b) cis-cup-tris(DPAF-C9) (cis-cup-2-C9) for comparison.
Molecules 24 03337 g002
Figure 3. Infrared spectra (KBr) of (a) C60, (b) C60(>DPAF-C9) (1-C9), (c) tris[C60(>DPAF-C9)]. and (d) tris[(DPAF-C2M)-C60(>DPAF-C9)] (4-C2M−9).
Figure 3. Infrared spectra (KBr) of (a) C60, (b) C60(>DPAF-C9) (1-C9), (c) tris[C60(>DPAF-C9)]. and (d) tris[(DPAF-C2M)-C60(>DPAF-C9)] (4-C2M−9).
Molecules 24 03337 g003
Figure 4. UV-vis spectra of (A) (a) tris(DPAF-C2M) (2-C2M), (b) tris(BrDPAF-C2M) (3-C2M), (c) cis-cup-tris[C60(>DPAF-C9)], (d) C60(>DPAF-C9) (1-C9), and (e) cis-cup-4-C2M–9, where (a), (b), and (e) were taken in CDCl3 and (c) and (d) were taken in toluene. (B) Fluorescence spectra of (a) tris(DPAF-C2M) (λex: 352 nm), (b) C60(>DPAF-C9) (λex: 406 nm), and (c) cis-cup-4-C2M–9 (λex: 410 nm). The concentration of all samples is 1.0 × 10−5 M.
Figure 4. UV-vis spectra of (A) (a) tris(DPAF-C2M) (2-C2M), (b) tris(BrDPAF-C2M) (3-C2M), (c) cis-cup-tris[C60(>DPAF-C9)], (d) C60(>DPAF-C9) (1-C9), and (e) cis-cup-4-C2M–9, where (a), (b), and (e) were taken in CDCl3 and (c) and (d) were taken in toluene. (B) Fluorescence spectra of (a) tris(DPAF-C2M) (λex: 352 nm), (b) C60(>DPAF-C9) (λex: 406 nm), and (c) cis-cup-4-C2M–9 (λex: 410 nm). The concentration of all samples is 1.0 × 10−5 M.
Molecules 24 03337 g004
Figure 5. Cyclic voltammograms (CV) of (a) C60(>t-Bu-malonate)2, (b) cis-cup-tris[(DPAF-C2M)-C60(>DPAF-C9)], and (c) C60(>DPAF-C9), where (A) displays the sequential redox cycles of each compound with the assignments and (B) shows the corresponding potential voltage values at either the peak maximum or minimum of each redox cycle. All solutions were in a concentration of 5.0 × 10−3 M in CH2Cl2 using (n-butyl)4N+-PF6 as the electrolyte (0.1 M), Pt as working and counter electrodes, and Ag/AgCl as the reference electrode at a scan rate of 10 mV/s.
Figure 5. Cyclic voltammograms (CV) of (a) C60(>t-Bu-malonate)2, (b) cis-cup-tris[(DPAF-C2M)-C60(>DPAF-C9)], and (c) C60(>DPAF-C9), where (A) displays the sequential redox cycles of each compound with the assignments and (B) shows the corresponding potential voltage values at either the peak maximum or minimum of each redox cycle. All solutions were in a concentration of 5.0 × 10−3 M in CH2Cl2 using (n-butyl)4N+-PF6 as the electrolyte (0.1 M), Pt as working and counter electrodes, and Ag/AgCl as the reference electrode at a scan rate of 10 mV/s.
Molecules 24 03337 g005
Figure 6. (a) UV-vis spectra of α,α’-(anthracene-9,10-diyl)bis(methylmalonic acid) (ABMA) and fluorescence (FL) emission spectrum of (b) ABMA, (c) cis-cup-tris[(DPAF-C2M)-C60(>DPAF-C9)] (cis-cup-4-C2M–9), and (c) a combination of ABMA and cis-cup-4-C2M–9 in a solvent mixture of DMF–CHCl3 (1:9, v/v) in a concentration of 10−6 M.
Figure 6. (a) UV-vis spectra of α,α’-(anthracene-9,10-diyl)bis(methylmalonic acid) (ABMA) and fluorescence (FL) emission spectrum of (b) ABMA, (c) cis-cup-tris[(DPAF-C2M)-C60(>DPAF-C9)] (cis-cup-4-C2M–9), and (c) a combination of ABMA and cis-cup-4-C2M–9 in a solvent mixture of DMF–CHCl3 (1:9, v/v) in a concentration of 10−6 M.
Molecules 24 03337 g006
Figure 7. Fluorescence (FL) emission spectra of (A) ABMA and (B) tetrafluorofluorescein-10′ (or 11′)-carboxylate (TFFC) to correlate directly the singlet oxygen (1O2) and superoxide radical (O2·) production efficiency, respectively, with (a) blank control and (b) a mixture of corresponding probe and cis-cup-4-C2M–9 samples in DMF–CHCl3 (1:9, v/v) at a concentration of 1.0 × 10−5 M using ABMA as the 1O2 trapping agent at λex 350 nm and λem 428 nm and DNBs-TFFC as the O2·-acceptor agent at λex 484 nm and λem 530 nm for detection with the irradiation of white light emitting diode (LED) light.
Figure 7. Fluorescence (FL) emission spectra of (A) ABMA and (B) tetrafluorofluorescein-10′ (or 11′)-carboxylate (TFFC) to correlate directly the singlet oxygen (1O2) and superoxide radical (O2·) production efficiency, respectively, with (a) blank control and (b) a mixture of corresponding probe and cis-cup-4-C2M–9 samples in DMF–CHCl3 (1:9, v/v) at a concentration of 1.0 × 10−5 M using ABMA as the 1O2 trapping agent at λex 350 nm and λem 428 nm and DNBs-TFFC as the O2·-acceptor agent at λex 484 nm and λem 530 nm for detection with the irradiation of white light emitting diode (LED) light.
Molecules 24 03337 g007

Share and Cite

MDPI and ACS Style

Yin, H.; Wang, M.; Tan, L.-S.; Chiang, L.Y. Synthesis and Intramolecular Energy- and Electron-Transfer of 3D-Conformeric Tris(fluorenyl-[60]fullerenylfluorene) Derivatives. Molecules 2019, 24, 3337. https://doi.org/10.3390/molecules24183337

AMA Style

Yin H, Wang M, Tan L-S, Chiang LY. Synthesis and Intramolecular Energy- and Electron-Transfer of 3D-Conformeric Tris(fluorenyl-[60]fullerenylfluorene) Derivatives. Molecules. 2019; 24(18):3337. https://doi.org/10.3390/molecules24183337

Chicago/Turabian Style

Yin, He, Min Wang, Loon-Seng Tan, and Long Y. Chiang. 2019. "Synthesis and Intramolecular Energy- and Electron-Transfer of 3D-Conformeric Tris(fluorenyl-[60]fullerenylfluorene) Derivatives" Molecules 24, no. 18: 3337. https://doi.org/10.3390/molecules24183337

Article Metrics

Back to TopTop