Functional Naphthalene Diimides: Synthesis, Properties, and

Aug 26, 2016 - Following her Ph.D., she undertook a higher degree in Education at Cambridge University (Sidney Sussex College), and postdoctoral studi...
10 downloads 27 Views 111MB Size
Review pubs.acs.org/CR

Functional Naphthalene Diimides: Synthesis, Properties, and Applications Mohammad Al Kobaisi,† Sidhanath V. Bhosale,‡ Kay Latham,† Aaron M. Raynor,† and Sheshanath V. Bhosale*,† †

School of Applied Sciences, RMIT University, GPO Box 2476, Melbourne, Victoria 3001, Australia Polymers and Functional Materials Division, CSIR-Indian Institute of Chemical Technology, Hyderabad, Telangana-500007, India



ABSTRACT: This comprehensive review surveys developments over the past decade in the field of naphthalene diimides (NDIs). It explores their application toward: supramolecular chemistry; sensors; host−guest complexes for molecular switching devices, such as catenanes and rotaxanes; ion-channels by ligand gating; gelators for sensing aromatic systems; catalysis through anion−π interactions; and NDI intercalations with DNA for medicinal applications. We have also explored new designs, synthesis, and progress in the field of core-substituted naphthalene diimides (cNDIs), and their implications in areas such as artificial photosynthesis and solar cell technology. Also presented are some interesting synthetic routes and procedures that can be used toward further development of NDI-bearing compounds for future applications. Finally, we conclude with our views on NDI chemistry for future research endeavors, and we outline what we believe are the key obstacles that need to be overcome for NDIs to see real world applications.

CONTENTS 1. Introduction 2. Synthetic Methods 2.1. Synthetic Strategies of Core-Substituted NDIs 3. Supramolecular Self-Assembly 3.1. Chemistry beyond the Molecule: Bioinspired Approach 3.2. Amino Acid and Peptide Derivatized NDIs 3.3. Gelation and Solvophobic Controlled Assembly 3.4. Self-Assembly of Core-Substituted NDIs 3.5. Self-Assembly through Donor−Acceptor CT Complexes 3.6. Metal-Ion Coordinated Self-Assembly 3.7. Self-Assembly through Molecular Recognition 3.8. Chiral Self-Assemblies 4. Molecular Sensors 4.1. Anion−π Interactions and Anion Sensors 4.2. Cation Sensors 4.3. pH Sensors 5. Host−Guest CT-Type Complexes 5.1. Catenanes and Rotaxanes 5.2. Pseudorotaxanes 5.3. Catenanes 6. Ion-Channels 7. Artificial Photosynthesis 8. Medicinal Application 9. Optical Properties and Optoelectronics 9.1. Small Optoelectronic Molecules © 2016 American Chemical Society

9.1.1. Passive Alkyl, Vinyl, Ethynyl, and Aryl Substitutions 9.1.2. Fluorinated Substituted NDI 9.1.3. Heterocyclic Fused and Core Expended NDIs 9.1.4. Ionic and Water-Soluble Functionalized NDIs 9.1.5. Donor−Acceptor Systems 9.2. NDI Based Polymers 9.3. Role of Spacers on NDI Optoelectronics 10. Catalysis 11. Conclusion and Outlook Author Information Corresponding Author Notes Biographies Acknowledgments References

11685 11687 11688 11691 11691 11694 11698 11700 11702 11706 11709 11713 11713 11714 11716 11718 11723 11724 11728 11728 11729 11731 11746 11753 11754

11754 11755 11759 11764 11766 11770 11775 11776 11778 11779 11779 11779 11779 11780 11780

1. INTRODUCTION Naphthalene diimide (1) (NDI) is the smallest homologue of the rylenediimides (RDI) (2), among which perylenediimide (PDI, n = 1, Figure 1) is no-doubt the most important.1−3 Both NDIs and RDIs possess high electron affinity, good charge carrier mobility, and excellent thermal and oxidative stability, making them promising candidates for organic electronics applications, photovoltaic devices, and flexible displays.2,4−6 On Received: February 28, 2016 Published: August 26, 2016 11685

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

(LUMO energy = −3.7 eV). Furthermore, the electronwithdrawing groups at the imide position cause a strong polarization of the π-systems, and the aromatic naphthalene core thus possesses a low π-electron density. Imide substitution has very little effect on the optical and electrochemical properties of NDIs, and they can be functionalized with a variety of groups at the imide position without changing the electronic properties of the π-scaffold in the development of aggregation behavior. Imide-functionalized NDIs have shown potential as radical anions with high conductivity,23 in artificial photosynthesis,24−29 in supramolecular self-assembled architectures,30−35 and in donor−acceptor systems.36,37 NDIs bearing amino acid and peptide derivatives have also been used to fabricate supramolecular self-assemblies from 1D to 3D nanostructures.38 In particular, NDIs bearing amino acids at the imide positions build-up into helical nanorods that can act as receptors for C60.39,40 Functionalization through the diimide nitrogens produces nonfluorescent or weakly fluorescent NDI analogues, which are capable of self-organization41 and have been incorporated into larger multicomponent assemblies through intercalation.42 π-Acidity of NDI, as a result of anion−π interaction, has also been reported,43,44 and it has been shown that, in the presence of NDI derivatives, these interactions play an important role in anion transport across lipid membranes. In the latter work it was argued that the rod-like oligo-NDI served as an anion slide, and that the anions at the NDI units are able to hop from one binding site to the next to pass through the channel.43,44 Matile and co-workers applied the strong π-acidity of the NDI moieties through NDI−anion interactions and their catalytic activity,41 and also in transport of anions across lipid membranes.43−45 In this regard, the Saha group have shown that electron-deficient NDIs interact strongly with basic anions, such as fluoride. Here, an electron-transfer (ET) process occurs between the anions and the electron-deficient NDI, which can take place thermally or be photoinduced (PET).46−48 In further developments, Iverson and co-workers studied the chargetransfer (CT) complex motif of electron-poor NDI derivatives, with electron-rich 1,5-dialkoxynaphthalene (DAN), to construct foldamers through CT interactions. Such weak CT interactions result in large molecules with defined nanostructures with a flexible network.49 The alternate stacked NDI acceptors and DAN donors have been used as templates for the synthesis of the respective catenanes,50−52 rotaxanes,53,54 synthetic ion channels,35 and more complex structures, such as knots.55 These alternate NDI and DAN derivatives, with amide groups at the imide position, evoke color enhancement of colorless NDI derivatives through CT complex formation.56,57 Taking advantage of this CT complex, the Shinkai group described preparation of NDI-containing organogels for colorimetric sensing of regio-isomeric dihydroxynaphthalenes.41 Until 2000, NDIs had only been studied with modification at the imide position, and no chemical modification of the four (2, 3, 6, and 7) available positions of the NDI core had been reported.58 This may be due to the Vollmann report in the early 1990s, which showed that core-substitution of NDIs with aryl functionalities did not change any of its properties.59 However, the situation changed over the past few years, and NDI derivatives now represent one of the most explored subclasses of the rylene family. In 2002, the Würthner group showed that by placing heteroatom (N, O, or S) donor substituents at the NDI core, producing core-substituted NDIs (cNDIs), all possible colors and optical properties,60 and wide-ranging

Figure 1. Aromatic diimide structures with potential photophysics, self-assembly, and functional properties.

the other hand, 1,4,5,8-naphthalenetetracarboxylic acid dianhydride (NDA) (3), which is the main precursor for the synthesis of a variety of NDI derivatives, has also been used as a precursor for the production of industrially relevant perinone pigments,7 and it can be readily functionalized, through the anhydride position, with arylamino or alkylamino groups. PDI (2)2,3,8 and the higher analogues, the rylene diimides,9−11 have been used for many years, due to their outstanding fluorescent properties.12 There have been many reviews of these materials, but the focus of these reviews has been somewhat different from that herein. Most have concentrated on their application, for example in fluorescence spectroscopy;13 organic electronic and photovoltaic devices;5,14 and metallo-supramolecular15 and supramolecular architectures,16 as well as their interaction with DNA and RNA.17 Furthermore, Würthner et al. recently detailed the synthesis and supramolecular assemblies of PDI dyes in depth.18 The main emphasis of this review is on the chemistry of NDIs and to give an update on their design, synthesis, properties, and applications. As such, PDI, core-enlarged systems, such as terrylene or quaterrylenebisimides, and other annulated scaffolds consisting of PDI subunits, will not be included in this discussion.18 NDIs are a neutral, planar, chemically robust, redox-active, electron-deficient class of aromatic compound, which usually have high melting points, and they have been used for a variety of applications, ranging from biomedicine to electronics.19 In the design of electronic conducting functional materials, NDIs are among the most versatile and fascinating class of aromatic molecules.20 Nevertheless, functionalization through core-substitution (2,6- or 2,3,6,7- to the core) (4) produces NDI analogues whose absorption and fluorescence properties are variable.21 A most important feature of the NDI chromophore is its mirror image fluorescence, which has a fluorescence quantum yield very close to unity.8−12 Indeed, high fluorescence quantum yields (∼0.9) are observed for cNDIs in all common organic solvents (aliphatic, aromatic, chlorinated, and dipolar solvents). Functionalization of NDI through the imide nitrogen atoms, exclusively, i.e. with N,N′-dioctyl-substitution, results in colorless solids, which in CH2Cl2 absorb in the UV region (i.e., wavelength smaller than 400 nm), and has a weak mirror image emission with a 7 nm Stokes shift. On the other hand, in toluene, excimer-like emissions have been reported, suggesting that ground-state aggregates are formed readily.22 NDI derivatives with two annulated electron-withdrawing imide groups possess a planar aromatic scaffold and are easily and reversibly reducible. Typically, NDI derivatives in CH2Cl2 show two reversible one-electron reductions: E1/2Red1 = −1.10 V and E1/2Red2 = −1.51, versus ferrocene/ferrocenium (Fc/Fc+) 11686

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

frontier molecular orbital levels,61 could be achieved, which aided the study of artificial photosynthetic processes.62 Since then, researchers have functionalized NDI in many ways, including alkylamino-, alkoxy-, sulfur-,60 aryl,63−65 thiophene,66 cyano,67 and metallo-bridging, resulting in intensely colorful, highly conductive functional materials, with significantly different photophysical properties than their counterparts without core-substitution.21 cNDI properties can thus be tuned over a wide wavelength range by varying the electrondonating ability of the core substituents.68−70 Thereafter, a number of groups used core-functionalized NDI components for the creation of supramolecular functional materials.63,71−77 NDI chemistry has transpired as a result of their desirable electronic and spectroscopic properties, over pyromellitic diimide (PMI) (5), and has better fabrication properties than the PDI dyes due to their enhanced solubility.18 There has also been growing interest to use cNDIs in biological and medical applications. This review provides a comprehensive survey of recent developments in the chemistry of naphthalene diimides (NDIs) and their applications in the fields of organic, biomolecular, optoelectronic, and supramolecular science. It commences with a comparison of structurally similar derivatives; their general use; methods of synthesis; electronic, spectroscopic, and chemical properties; and their usefulness in optical devices. This is followed by discussion of their application to the fields of material science and molecular sensors and of the structure− function relationship. Also discussed are the development of molecular systems and supramolecular arrays, and recent examples of donor−acceptor charge transfer (CT) complex systems within the areas of host−guest chemistry, including foldamers, catanenes, rotaxanes, and synthetic ion channels. Finally, we explore the developments of NDIs and coresubstituted NDIs in the field of artificial photosynthesis, i.e. energy and electron transduction systems using the covalent and supramolecular self-assembly approach. We conclude with our views and a brief summary and outlook of future endeavors in NDI chemistry.

Figure 2. Synthesis pathways of NDI derivatives.

2. SYNTHETIC METHODS Naphthalene diimides (NDI) have made significant contributions to the evaluation of design principles because of their ease of synthesis, and they have also attracted much attention due to their tendency to form n-type semiconductor materials.20 Typically, the synthesis of symmetric NDI compounds (Figure 2) is a simple and efficient one-step procedure in which commercially available 1,4,5,8-naphthalenetetracarboxylic acid dianhydride, i.e. NDA, is condensed with the appropriate primary amine (a variety of amines can be used) in N,N′dimethylformamide (DMF). A variety of functional groups can be introduced at the imide-position, using this procedure. Furthermore, core-functionalization at the 2,6- and 2,3,6,7positions by N, O, or S donor substitution has led to very different photophysical properties than those of their coreunsubstituted counterparts. A recent literature survey revealed that there have been numerous applications of NDI molecules, from sensors to solar cells. This area has been the subject of reviews by us,1,58 by Matile et al.,61 and very recently, by Würthner et al., who gave in-depth detail on novel synthetic pathways.21 Thus, in this section only a few interesting examples will be provided to illustrate the novel synthesis of NDIs (6−60). In general, symmetrical NDI derivatives are easily prepared by a simple, one-step procedure, condensing

NDA with the appropriate primary amine, in a high boiling solvent such as isopropanol or DMF, for 12−24 h at 70−110 °C.1−15 However, the synthesis of unsymmetrical NDIs, with two different amines at the imide positions, results in a mixture of products, which are difficult to separate and purify. However, a report by the Ghadhiri group showed that unsymmetrical NDIs can be prepared, and in very good yield, by use of a solid support-based synthetic method.78 This procedure involves hydrolyzing NDA to the tetracarboxylic acid form with KOH, followed by reacidification to pH 6.2 with H3PO4, to form the monoanhydride. Following addition of 1 equiv of a primary amine to this reaction mixture, and heating at reflux overnight, the naphthalene monoimide (6) is produced, in good yield and purity. In the second step, the monoimide NDI is reacted with primary amines, or with a cyclic peptide directly, without isolating the anhydride intermediate, with refluxing in DMF producing the unsymmetrical NDIs (7−11). An improved and simpler synthetic protocol was developed by the Sanders group, where a microwave-assisted synthetic procedure was employed for the synthesis of structurally diverse NDIs, in very high yield, and within 1 to 5 min (12 and 13).79 This approach gives flexibility to use a wide variety of imides, and also led to the conclusion that the transformation is compatible with acidlabile protecting groups such as protected α-amino acids. The 11687

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 3. Examples of the synthetic pathways of 2-substituted NDI derivatives.

the natural product, asparagusic acid. The initiators and propagators consist of a series of complex NDIs prepared from NDA. The main feature is the inclusion of four germinal diphosphonates for tetravalent anchoring on the indium tin oxide (ITO) surface. The syntheses include a series of microwave-assisted imidation reactions, followed by a series of peptide coupling reactions, to build-up the side chains of the central NDI unit. With the building blocks in hand, SOSIP is carried out in a solid-phase fashion on the ITO surface by immersing the electrodes in a 3 mM solution of the initiator in DMSO for 2 days. The propagators were then introduced into the system in a similar fashion at their critical SOSIP concentration (CSOSIP). The polymers were characterized by UV−vis absorption spectroscopy.

synthesis is simple and requires only, on average, minutes in a microwave reactor in the presence of DMF and TEA. Generally most products can be obtained in pure form after simple workup, and further purification was only required in a few cases. No side reactions due to either loss of the protecting group or self-condensation of the α-amino acids were observed. They have also examined the possibility of racemization by preparing the symmetrical NDIs from racemic H-Cys(Trt)-OH and (R)-H-L-Cys(Trt)-OH. It was clearly shown that formation of a diastereomeric mixture and (R,R)-NDI does not occur; thus, the reaction proceeds with no racemization at the αcenter. The utility of this reaction was then further examined to include unsymmetrical NDIs. These were subsequently obtained in high yields with very little contamination of the symmetrical by product. Using similar strategies, Pantos and coworkers reported the stepwise and high-yielding synthesis of symmetric and asymmetrical NDIs, at low temperature (14, 15). In their study they also investigated the influence of factors in unsymmetrical NDI synthesis, for example, steric effects, solubilities of amines, and noncovalent interaction between reagents.80 These procedures provide control over the stepwise condensation of primary amines with NDA, to result in unsymmetrical NDI derivatives in excellent yield. A very interesting example by the Stoddart group reported an enantiomeric pair of doubly bridged NDI cyclophanes (16), and the characterization of four of their electronic states, namely (1) the ground state, (2) the exciton coupled singlet excited state, (3) the radical anion with strong through-space interactions between the redox-active NDI molecules, and (4) the diamagnetic diradical dianion, using UV/vis/NIR, EPR, and ENDOR spectroscopies, in addition to X-ray crystallography. Despite the unfavorable Coulombic repulsion, the singlet diradical dianion dimer of NDI shows a more pronounced intramolecular π−π stacking interaction when compared with its neutral analogue.81−83 Höger and Rauch have developed a new and efficient synthetic route for the core-amino substitution of NDIs (18, 19) and PDIs, under mild conditions, by reacting alkyl-substituted NDI (17) with primary or secondary amines, Cu(II) salts as catalysts, and air as an oxidant.82 The developed method is easy to use, very fast, and neither air nor moisture sensitive, and the final product yield is over 80%, all by simply stirring the reaction mixture at room temperature. A recent review article by the Matile group discusses the efforts and achievements made by organic chemists in the area of materials science.84 The discussion outlines the preparation of supramolecular n/p-heterojunctions with oriented multicomponent/color antiparallel redox gradients abbreviated to OMARG-SHJ. The unique architecture of OMARG-SHJ is created by a self-organizing surface initiated polymerization (SOSIP), controlled by five unique building blocks of organic initiators and propagators. The complete synthesis of each compound is discussed in detail, including the total synthesis of

2.1. Synthetic Strategies of Core-Substituted NDIs

While affecting the solubility and molecular organization significantly, imide-functionalization of NDI has little effect on the molecular-level electronic and optical properties. However, the presence of one or more substituents at positions 2, 3, 6, and 7 of core-substituted NDIs (cNDIs) affects both their electronic and optical properties.57,60 cNDIs have been reported in the literature since the early 1930s58 but have been produced with simplicity, only in the latter part of the 20th century. In order to utilize NDIs as dyes for applications in electronics, increased electronic coupling between the NDI core and the appended electroactive elements is required, while aggregation effects, orthogonal to the aromatic plane, should be limited to minimize optical “short-circuiting” based on crystallization. This has been achieved by core-substitution of NDI with complementary electro-active elements covalently bound to the naphthyl core of the NDI.13,14 In this regard, a number of cNDI derivatives, with aryl, thiophene, and cyano groups, have been employed. However, expansion of the πsystem along the lateral position of cNDIs has been limited due to synthetic difficulties. cNDIs are usually prepared from the respective halogenated NDA (chloro or bromo), but preparation of chloro-NDA requires harsh conditions. However, bromo-core-substitution is much easier, and bromination can be controlled to produce mono (21), di- (22), and tetrabrominated (NDA: 20, NDI: 23) forms.22,76 Since this time, there has been considerable progress in the synthesis and in the range of application of cNDIs. To this end, we recommend that interested readers study the excellent review of Würthner et al.21 Leng and co-workers synthesized a series of monolateral and bilateral sulfur-heterocycles fused cNDIs, starting from mono(21) and dibromo- (22) cNDIs via nucleophilic aromatic substitution reaction, followed by oxidative aromatization (25, 27), as shown in Figures 3 and 4.85 Furthermore, their electrochemical and optical properties were investigated, and they revealed that the monolateral cNDI derivatives possessed higher thermal stability, compared to their bilateral counter11688

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

polymers (NDI-alt-BBO (34) and NDI-r-OPV/BBO (35)). Bottom gate−top contact OFETs of NDI-alt-OPV exhibit ambipolar charge transport with average electron and hole mobilities of 3.09 × 10−3 cm2/V·s and 2.1 × 10−3 cm2/V·s, respectively (Figure 4). On the other hand, the random copolymer incorporating both OPV and BBO units, i.e. NDI-rOPV/BBO, showed dominant n-type charge transport with moderate electron mobility 4 × 10−4 cm2/V·s. Thus, the present work highlights the structure−property relationship and the electronic tunability, based on NDI-based polymers, which are required for ambipolar transistors. Figure 5 illustrates the synthetic pathways of 2,3,6,7substituted NDI derivatives (36−60), and the controlled orientation of isomers. A very simple protocol for the synthesis of six-membered NDI heterocyclic acene diimides (36, 37), by condensation of tetrabromo-NDI with ortho-phenylenediamine, 1,2-benzenedithiol, and 2-aminothiophenol was developed by Li et al.87 Typically, tetrabromo-NDI was treated with 1,2-benzenedithiol in DMF, in the presence of K2CO3, and heated at 100 °C for 1 h, to form the symmetrical acene diimide in almost quantitative yield (96%). However, when orthophenylenediamine was used, even with a high excess of reagent and extended reaction times (from 1 h to 2 days), only monocondensation was observed. This unexpected selectivity provided access to a series of asymmetrical, red to NIR absorption, hydroazaacene derivatives synthesized by Cai et al., and which exhibited well-defined J-aggregation behavior both in hydrocarbon solution (hexane or octane), as well as in thin films of cNDIs (38−40).88 The J-aggregation could be visualized by observing the characteristic colorless solution formed in a hydrocarbon solvent compared to the green solution obtained when chloroform was used. Taking advantage of asymmetrical substitution, Li et al. synthesized a series of heterocyclic acenes (41−45),87 and modification of the NDI core resulted in a change in the electronic properties of the materials from n-type to p-type semiconductors. Further investigation into the electronic properties of these materials also uncovered their potential as NIR dyes and potential electron donors following the introduction of NH bridges into the lateral position. To elaborate on these results, the semiconducting properties were also investigated in thin-film transistors (TFTs), which after annealing at 140 °C, exhibited a high hole mobility of 0.02 cm2/V·s for the asymmetrical acenediimide.6 Li et al. reported a pair of diazapentacenediimide derivatives, constructed around the NDI core (46, 47).89 Dihydro diazapentacenediimide derivatives were produced by the double cross-coupling of zirconacyclopentadienes with unilateral Br4−NDI, followed by dehydrogenate aromatization with lead dioxide (PbO2). Typically, the nucleophilic addition of o-phenylenediamine with Br4−NDI was carried out to form the monosubstituted product in 90% yield. This product was then used in a Cu(I) chloride catalyzed cross-coupling reaction with zirconacyclopentadiene reagents prepared in situ from alkynes and zirconocene dichloride (Cl2ZrCp2). Two separate reactions were performed to generate the dihydrodiazapentacenediimide derivative (46) with yields of 24%. These compounds were then oxidized with PbO2 to form the dihydrodiazapentacenediimide derivatives (47) in high yield. DDQ oxidation of the cyclohexyl ring was attempted, but no reaction occurred even after 5 days at 110 °C. The electronic properties of these compounds were then studied via cyclic voltammetry (CV) to reveal that both compounds exhibit reduction and oxidation waves, and thus show promise as

Figure 4. Synthesis pathways of 2,6-substituted NDI derivatives.

parts, and lower LUMO i.e. −4.0 eV. These results clearly show that these new cNDI derivatives may be useful for n-type organic semiconductors. Fukutomi et al. explored the expansion of core-substituted NDIs by developing a procedure for the synthesis of naphthodithiophenediimides (NDTIs) (29).86 After many attempts, the most efficient route was to start from 3,7-dibromo-NDI (22), which readily underwent a Stille coupling, to generate the bis(TMS)ethynyl intermediate (28). This precursor could then be cyclized in the presence of sodium sulfidenonahydrate (Na2S·9H2O) in ethanol, to produce 29 in 15% yield. Core-substitution with amino, sulfur, and oxygen functionality has been very well demonstrated.1,21,58 However, carbon− carbon core-substitution was first designed and carried out by Kolhe et al., where a series of alternating donor (D)−acceptor (A) copolymers, based on NDI as the acceptor and oligo(pphenylenevinylene) (OPV) (30) or benzobisoxazole (BBO) (31) as the strong or weak donor, were synthesized by employing Suzuki coupling and Horner−Wadsworth−Emmons polymerization with bis-aldehyde-NDI (32), respectively.81 The effect of donor strength on the photophysical, electrochemical, and semiconducting properties of the polymers was also investigated. NDI-alt-OPV (33) has both LUMO (∼0.2 eV) and HOMO (∼0.5 eV) energy levels, as compared to two other 11689

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 5. Examples of the synthesis pathways of 2,3,6,7-substituted NDI derivatives, and the control of orientation isomers. 11690

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

electron-donor materials. Contrary to this, compound (47) displayed only reduction waves, indicating potential as an electron-accepting material. All the symmetric and asymmetric diimide (36−47) derivatives were found to be very soluble in common organic solvents, such as chloroform, tetrahydrofuran, and toluene. The majority of acene derivatives are prone to atmospheric instability, such that when nitrogen atoms are incorporated into the acene skeleton to form azaacene derivatives, the resulting compounds are less susceptible to degradation through oxidation or dimerization. Another approach for the core extension of NDIs was reported by Suraru et al. This method, proceeding via the attachment of indole rings to the core, produced carbazolo[2,3-b]carbazole naphthalene diimides (CbNDIs) in two steps (48, 49).90 In the first step, 2,7-diphenylamino-3,6-dibromo-NDIs was prepared by reacting tetrabromo-NDIs with the respective aniline (approximately 90 equiv), in either chloroform or toluene, to produce yields of up to 87%. However, when electron-poor anilines such as 3,5-bis(trifluoromethyl)aniline were used, prolonged reaction times (7 days) in toluene and higher temperatures (70 °C) were required, to achieve practical yields. In the second step, cyclization of these diamine dibromo-NDIs was carried out with Pd(OAc)2 as the catalyst, in the presence of K2CO3 and with DMF as the solvent. High temperatures were required to achieve cyclization (100 °C) and produced only moderate yields, ranging from 28 to 54% for the isolated product. A new NDI derivative with a 4-amino group at the core of the NDI (4NH2−NDI) (51) was synthesized via nucleophilic substitution of 4Br-NDI 23 with benzophenone imine to give compound 50 in 59% yield. Subsequently, hydrolysis under mild conditions afforded compound 51 in 42% yield. Furthermore, their organic field effect transistor (OFET) properties were investigated.91 Interestingly, introduction of amino groups to the core of the NDI gives a high HOMO energy (−4.87 eV) and behaves as a p-type semiconductor with a rather low hole mobility of 8.2 × 10−5 cm2/V·s. However, the hole mobility is enhanced significantly with doping of core-unsubstituted N,N′-dihexyl-naphthalene diimide (DHNDI). Nevertheless, thin films of a blend of 4NH2−NDI and DHNDI (with a ratio 1:3) exhibit ambipolar behavior, based on the respective transfer and output characteristics measured in an inert atmosphere, and hole and electron mobilities of 0.01 cm2/V·s and 0.001 cm2/V·s, respectively. Asymmetric anti- and syn-isomers and quinonoid form heterohexacene diimides (52−60)92 containing NH and O/S were synthesized by condensation of 4Br-NDI 23with 2aminothiophenol or 2-aminophenol followed by oxidation using lead oxide. During synthesis of heterohexacene diimides, they observed the formation of two different regioisomers (anti- and syn-), but only anti-isomers could be oxidized into the quinonoid products. Interestingly, they have observed that syn-/anti- can be lowered at increased reaction temperature such that the reaction was conducted at 90 °C in DMF, and the ratio of syn-1/anti-1 dropped to ∼1.5:1. When other solvents were used, this also led to lowered selectivity (syn-1:anti-1 = ∼ 1:2.3 in CHCl3 at 60 °C and ∼1:1.8 in toluene at 110 °C). Importantly, both the stable quinonoid diimides displaying low LUMO levels at less than −4.1 eV are obtained via oxidation of the anti- isomers. Thus, the method described controlled synthesis and, in reducing the isolated quinoidal molecules back to dihydro-forms, offers a route to pure anti- isomers. The Bonifazi group developed acid-mediated, simple protocols for the rational synthesis of AB-type cNDIs, in very high yields.93

Typically, tetraethyl 2,6-diethoxynaphthalene-1,4,5,8-tetracarboxylate was prepared in two steps: first, bromination of commercial NDA with dimethyl dibromohydantoin, followed by alkylation with Me2SO4 to produce tetramethyl 2,6dibromonaphthalene tetracarboxylate (22), and then, in a second step, reaction of dibromonaphthalene with sodium ethoxide followed by acid-mediated transformation into the asymmetric core-substituted naphthalene anhydride ester (cNAE) in quantitative yield. Finally, cNAE was converted into hetero-N-substituted cNDI through sequential condensation reactions, in the presence of the appropriate amine. Recently, Maniam et al. described an unusual oxidation of NDIs under mild oxidizing conditions, using ruthenium(III) chloride and sodium periodate, to afford 1,4-diones in good yield.94 In another report, Bijak et al. synthesized five- and six- membered azomethine-naphthalene (AzNDIs) diimides via a condensation reaction of diamines containing an NDI core and two aldehydes: 4-(heptadecafluoroundecyloxy)benzaldehyde and 4-octadecyloxybenzaldehyde.95 All of the azomethine-diimides exhibited LUMO and HOMO levels in the ranges −3.87 to −4.2 eV and −5.56 to −6.02 eV, respectively. Matsidik et al. reported the role of aromatic solvents in polycondensation via C−H activation of the NDI core with thiophene. These results will be very useful for future development of n-type semiconducting materials in the future. There is interest in ntype π-conjugated NDI-based polymers, due to its ever growing and excellent organic field-effect transistor (OFET) behavior and decent electron mobility (0.85−3.2 cm2/V s); therefore, a simple protocol is required in order to synthesize these polymers.96 A very interesting synthesis example was reported by the Chen group, who reported a simple method for the synthesis of pillar[6]arenes in modest yield (30%), by employing NDI as a template, in which the electron-deficient NDI and a 1,4-diisobutoxy-2,5-bis(methoxymethyl) benzene pillararene precursor interact strongly via donor−acceptor interactions.97

3. SUPRAMOLECULAR SELF-ASSEMBLY Supramolecular chemistry, or “chemistry beyond the molecule”, mostly focuses on the study of molecular recognition, and on high-order assemblies formed by noncovalent interactions.98 Self-assembly and the self-organization of molecules play important roles in the construction of thermodynamically stable structures at both the cellular and subcellular levels, within nanometer to millimeter dimensions, and they utilize a “bottom-up” or “bioinspired” approach. Supramolecular systems are made from building blocks that are linked together by noncovalent interactions, which can show stimuli-responsive behavior.99 3.1. Chemistry beyond the Molecule: Bioinspired Approach

Nature uses supramolecular self-assembly approaches to construct complex molecular architectures, which are crucial for sustaining life. For example, the bacterial center for lightinduced charge separation contains several membrane proteins, which fixate one bacteriochlorophyll molecule, two bacteriopheophytins, and one quinone, along an axis through the membranes. Such “machinery of life” developed biologically over a long-term evolution process. In order to mimic the molecular architectures of such well-designed assemblies, one must consider with equal importance both covalent and noncovalent chemistry in the creative process. As such, the strategy of coupling traditional synthetic chemistry with 11691

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 6. Various self-assembled nanostructures produced by aggregates of NDI derivatives: (A) the π-sheet self-assembly nanobelts formed by ntype dilysine peptide (Fmoc-KK(NDI)) functionalized at the ε-amino position with an NDI. Reproduced with permission from ref 147. Copyright 2010 the Royal Society of Chemistry. (B) Left-handed helical nanofibers produced by AcKK(NDI)-NH2 dipeptide-NDI. Reproduced from ref 148. Copyright 2009 American Chemical Society. (C) Nanotubes assembled from NDI−lysine amphiphiles at aqueous pH 12 and (D) pH 7. Reproduced with permission from ref 149. Copyright 2011 Wiley-VCH. (E) Left-handed twisted nanoribbons produced by AcK(NDI)K-NH2 dipeptide-NDI. Reproduced from ref 148. Copyright 2009 American Chemical Society. (F) Helical tapes self-assembly produced by NDI-lysine-NH2 amphiphiles at aqueous pH 7. Reproduced with permission from ref 149. Copyright 2011 Wiley-VCH. (G) Vesicular self-assembly of nonionic bolaamphiphile NDI-amide-Ph-PEG in aqueous medium. Reproduced with permission from ref 150. Copyright 2012 Wiley-VCH.

supramolecular chemistry, which uses weak interactions, such as hydrogen bonding, π-stacking, metal coordination bonding, and van der Waals forces, has proved particularly successful for the design and reproduction of a variety of artificial architectures of different sizes, shapes, and functions, similar to those of the working molecular apparatus.100−104 Supramolecular chemistry began as early as the 1960s, with the discovery of “crown ethers”, “cryptands”, and “spherands”, by Pedersen,105 Lehn,106 and Cram,107 respectively. These complementary molecules were able to recognize each other

and their components and to self-assemble through noncovalent interactions. A literature survey revealed that there are many types of self-assembly ranging in dimensions from nanoto micrometric, to macroscopic in size (visible to the naked eye), and including constructions such as coordination nanotubes,108 nanoballs,109 nanoboxes,110 and self-assembled cages.111 These self-assembled architectures provide spaces/ cavities, which can be used for novel physical chemistry, for example color sensing following guest inclusion.112 For this approach, various small molecules have been utilized, e.g. 11692

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 7. (A) Tapping-mode AFM image on freshly cleaved mica. Inset: a single fiber showing left-handed helicity along the long axis of the nanofiber; section analysis between two crosses in the AFM image showing regular helical pitch, 68, AcKK-NDI-NH2. (B) Ribbons showing lefthanded twists imaged using tapping mode AFM of 68, Ac-K(NDI)K-NH2. Inset: section analysis between two crosses in the AFM image showing the smooth surface of the ribbons. Scale bar: 100 nm. Reproduced from ref 148. Copyright 2009 American Chemical Society. (C) AFM image of the self-assembly of dipeptide Fmoc-KK(NDI)-NH2 in water. Reproduced with permission from ref 147. Copyright 2010 the Royal Society of Chemistry.

porphyrins113 and oligo(p-phenylenevinylene) (OPV) derivatives,114 and several have exhibited molecular-level machine-like functions.112−114 However, among the different classes of molecules examined, the self-assembly of linear small πconjugated systems is at center stage, due to their favorable optical and electronic properties and hence their potential utility in organic electronic devices.115−122 The formation of capsules/cages through self-assembled processes has been investigated by several groups. To give a brief snapshot of the work in this area, the Fujita group have performed revolutionary research on the self-assembly of coordination capsules using square-cornered Pd(II) complexes with coligands,123 which were further used for coating formation with multivalent proteins.124 The joint work of the Crego-Calama and Reinhoudt groups is also noteworthy, describing the formation of calix[4]arene-based molecular boxes and their guest binding capabilities,125 while Atwood used hydrogen-bonding, rather than coordination, for the preparation of supramolecular capsules. However, one of the best examples of the preparation of a chiral spherical molecular architecture, held together by 60 hydrogen-bonds, consisted of six calix[4]resorcinarenes and eight water molecules.126,127 The Rebek group prepared such a molecular capsule through the self-assembly of cavitands,128 and went on to demonstrate an unusual heteroguest inclusion situationthe preferential accommodation of one molecule each of benzene and pxylene.129 And finally, one must not forget the development of catenanes and rotaxanes, which are exceptional because of their well-defined, discrete supramolecular structure, and the diversity of their applications.130 This is a fine example where supramolecular self-assembly was used for creative functionality. Naphthelene diimides (NDIs) have made useful contributions to the evaluation of design principles because of their ease of preparation and n-type electron-accepting properties, along with their neutral, planar, and chemically robust nature. They are also redox-active with high melting points, and hence have similar properties to the naturally occurring acceptors of bacterial photosynthetic reaction centers.1,21,58,61 From this viewpoint, NDIs are one of the preferred classes of molecules. The optical and electronic properties of these molecules strongly depend upon their structural features and hence can be modulated by substitution at either of the diimide positions,

and/or directly onto the core of the NDI. An alternative approach to the variation of the electronic properties is by inducing intermolecular interaction using noncovalent forces such as H-bonding, coordination, and π-stacking, which has been widely exploited for the self-assembly of NDIs and for their application in forming various self-assembled architectures, e.g. cages, catenanes and rotaxanes, nanotubes, nanowires, sensors, ion channels, and applications. Core-unsubstituted NDI derivatives have been used widely for self-assembly, aggregation, induction of chirality, molecular recognition, molecular machines, molecular sensors, gas absorption, nanoreactors, chemical catalysis, drug delivery, optoelectronics, and chemosensors, to name a few.23,131−133 Supramolecular research is thus often cross-disciplinary, encompassing organic, physical, and coordination- and polymer-chemistry, materials science, biological science, engineering, etc.134 For the development of functional devices, π−π interactions (aromaticity) and solvophobicity are important for the production of supramolecular nanostructures.135 In this regard, well-defined nanostructures such as nanowires, vesicles, nanobelts, and gels have been produced from large macrocyclic polyaromatics, e.g. hexabenzocoronene and PDI.136−144 However, among those aromatic molecules that have found utility, and especially in the design of conducting materials, NDIs have attracted much attention, due to their interesting electro-optical properties and their tendency to form n-type semiconductor materials.24 Importantly, NDIs contain an aromatic core, which exhibits stacking in the solid-state via π−π stabilization, and four polar carboxyl groups, and hence are usually soluble in low polarity (DCM, CHCl3, toluene) and polar aprotic (acetonitrile, DMF) solvents, depending on the substituent on the imide. This phenomenon is useful for producing continuous stacks in supramolecular applications, but it can also be a major hindrance if the solubility is poor. Thus, the solubility of NDIs mainly depends on the substituents on the imide position, and generally NDIs with long and bulky aliphatic substituents have better solubilities.145 This feature is also useful for aggregation of NDI in solution and in the solid-phase.146 The development of strategies to introduce π-electronic nanostructures from n-type NDI is a topic of current interest to the authors. In this section, the supramolecular self-assembly, ranging from the nanometer to the micrometer scale, of small, 11693

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 8. (A−C) 62 derivatives. Reproduced with permission from ref 149. Copyright 2011 Wiley-VCH. (D−G) 144 derivatives. Reproduced with permission from ref 154. Copyright 2013 the Royal Society of Chemistry. (H−K) 143 derivatives, Reproduced with permission from ref 155. Copyright 2014 Wiley-VCH. (L and M) 76 DD and LL isomers. Reproduced with permission from ref 156. Copyright 2011 Wiley-VCH. (N and O) 77 and 78. Reproduced with permission from ref 157. Copyright 2011 the Royal Society of Chemistry. (P−T) 84 in various ethyl acetate to methanol ratios and concentrations. Reproduced with permission from ref 158. Copyright 2014 the Royal Society of Chemistry. (U and V) 69 DL and LL and DD derivatives. Reproduced with permission from ref 159. Copyright 2012 Wiley-VCH. (W) 92. Reproduced with permission from ref 160. Copyright 2015 the Royal Society of Chemistry. (X) NDI 529. Reproduced with permission from ref 161. Copyright 2015 the Royal Society of Chemistry.

aromatic π-conjugated NDI (61−190) functional derivatives is discussed.

of the NDI. Examples of the outcomes are discussed below, and in particular those that have led to reasonably robust design strategies for particular functional structures. A simple method for fabricating controlled, well-defined 1D nanostructures, i.e. either helical nanofibers or twisted nanoribbons, was demonstrated by the Parquette group from dilysine peptides f unctionalized at the imide position of the NDI

3.2. Amino Acid and Peptide Derivatized NDIs

There has been considerable, and diverse, work carried out in this area, including the impact of aliphatic vs aromatic, straight chain (including chain length) vs branched, and chiral vs achiral amino acids and peptides, and also the position of substitution 11694

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 9. NDI-peptide and other symmetric and asymmetric N-substituted NDI cores resulting in supramolecular self-assembly.

chromophore (Figure 6).147−150,113−115 Transmission electron microscopy (TEM) and atomic force microscopy (AFM) revealed that dipeptide-NDI (68, AcKK-NDI-NH2) forms lef thanded helical, micrometer-long nanof ibers with a uniform diameter (10 ± 1 nm), as shown in Figure 7.148 On the other hand, dipeptide-NDI, in which the position of the free amine groups was reversed, appeared as twisted nanoribbons (68, Ac-K(NDI)K-NH2) 20−70 nm in width (Figure 6). βSheet assembly via hydrogen-bonding was observed between the peptide backbones, where a delicate balance between electrostatic repulsion and hydrophobic interactions played an important role, along with π−π association of the NDI core in

water. Furthermore, when Fmoc dipeptide was introduced at the imide position (68, Fmoc-KK(NDI)-NH2), gelation occurred.147 TEM and AFM images suggested that the nanostructures of the gel were composed of antiparallel βsheets that stacked into dimers, and are further assembled into flat, micrometer-long nanobelts with regular heights of 3.5 nm (Figure 7(C)). Pandeeswar et al.151 gave further insight into the relationship and correlation that exists between architecture and conductivity of n-type organic semiconductors. Using NDI with peptide substituents, to induce differences in their morphology resulting from the chiral nature of the amino-acid building 11695

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

ethanol mixture with EtOAc, to produce flower-like and sheetlike structures, respectively. These results demonstrated that assembly could be controlled by mixing poor and good solvents; thus, this method can be used for the development of nanotechnological applications. An example of excellent work on chirality transcription, and induction of chirality, was conducted by the Govindaraju group, who successfully demonstrated the above by employing homochiral, heterochiral, and achiral peptide conjugates of NDI derivatives (69, R1 = H, R2 = CH2Ph).159 The enhancement of chirality was demonstrated by temperature-dependent CD data, which showed high thermal stability of the peptide based chiral supramolecular assembly. Both homochiral forms (LL, DD) formed one-dimensional belt-like superstructures of >50 mm in length with helical signatures lef t and right, respectively (Figure 8(V)). However, a mixture of DL assembled into particular aggregates (Figure 8(U)). Temperature-dependent CD data showed high thermal stability of the peptide auxiliary-driven chiral supramolecular assembly of NDI conjugates. Moreover, an interesting phenomenon, coined as retentive helical memory, was reported. Remarkably, the homochiral LL and DD showed 1D molecular organization, unlike the zero-dimensional organization of heterochiral LD and DL. These observations, in principle, could support the mechanisms of both spontaneous deracemization and enhancement pathways for biological homochirality. This work is significant from a fundamental science perspective, as well as for potential applications of NDI−peptide conjugates in chiral optoelectronics, biomaterials, and chiral technology.162 These studies were further extended, by other groups, by synthesizing NDI derivatives functionalized using various amino acids (Figure 9 (61−94)). Structure−property correlation in the self-assembly process of this group of compounds was further studied by the Govindaraju group and categorized in 0D (spheres), 1D (fibers), and 2D (sheets) architectures, showing how minute structural mutations can result in significant structural and electron mobility changes. Examples, and their applications, include the following: Pascu et al., who presented a new route to functionalized single-walled carbon nanotubes (SWNTs),163 via noncovalent interactions and charge-transfer interactions, leading to dispersion of the complex in widely accessible organic solvents such as CHCl3, DMSO, etc. (75).164 Another example by the Demets’ group presented pioneering work on the self-assembled NDI derivative (N,N′-(ethyl-N″,N″,N″dimethyloctadecane ammonium)-1,4,5,8-naphthalenediimide bromide) (82), functionalized at the imide position, which self-organized over electrodes. These molecular materials were shown to be very efficient toward electrochemical photoreduction of oxygen under visible light.165 Billeci et al.166 studied the self-assembly of fluorescent diimidazolium salts in search of new fluorescent organic salts to be used in biomedical and electrochemical fields. They designed a series of N,N′bis(1-alkyl-3-propylimidazolium)NDI diiodides, with consideration of the alkyl chain length and alkyl chain fluorination, which give rise to the formation of H-type aggregates in THF and DMF solution, and in the solid state. The influence of endcapped alkyl chain lengths on supramolecular hydrogelation of NDI-capped dipeptides was studied systematically by the Lin group, with particular emphasis on the intermolecular interactions of hydrogelators in both the gel and solution state (63).167 Their key findings illustrated the dependence of the gelation pH on chain length and revealed an inverse linear relationship over a broad range of pH from basic to acidic

blocks, large variances in structure and performance were achieved. Such that one-dimensional, achiral derivatives formed long f lat ribbon-like structures that overlapped one another, whereas 2D, chiral auxiliaries resulted in f lat sheets with minimal edge-to-edge overlap. It was also observed that the electron mobilities of 1D materials were higher than those of their 2D counterparts, at 3.5 and 1.6 × 10−6 cm2/V·s, respectively. Further work involving the impact of lysine on self-assembly was conducted by the Parquette group, who later demonstrated that lysine-based NDI bola-amphiphiles (74) first assemble into rings, and that these rings further aggregate into micrometer-long nanotubes with uniform diameters (12 ± 1 nm), and a wallthickness of approximately 2.5 ± 0.5 nm,152 while lysine-based NDI amphiphiles (62) assemble into 1D multilamellar nanotubes by the rolling of bilayer ribbons in aqueous media. TEM images of (62) revealed the formation of nanotubes with uniform diameters of 14 ± 1 nm (Figure 8(A)).149 A terthiophene(3T)-NDI f unctionalized dipeptide (68, R = terthiophene) self-assembled into 1D nanostructures through β-sheet interactions and extensive π−π stacking (Figure 8(B)). Interestingly, this NDI-dipeptide derivative self-assembled into 1D twisted f ibers in organic solvents and formed a self-supporting organo-gel at low concentrations in CH2Cl2.153 The Govindaraju group reported important work on supramolecularly engineered organization, using aromatic amino acids, in particular L- and D-phenylalanine methylesterappended NDIs (76). Field-emission scanning electron microscopy (FESEM) revealed formation of free floating nanosheets, of micrometre-size lateral dimension, from 90% aqueous acetonitrile solution. These nanosheets exhibited remarkable conductivity at ∼1.6 S/cm, and by controlling solvophobic effects, the derivatives could assemble into 0.1−1.5 attoliter (10 −18 L) nanocups, meso-cups, and bowl-like architectures from mixed, chlorinated solvents (Figure 8(L,M)).156 For example, nanocups were obtained from 50% (v/v) CHCl3/MeOH, meso-cups from 10% (v/v) CHCl3/ MeOH, and bowls from 10% (v/v) CCl4/MeOH, respectively. The group later built on this work and looked into extended aromatic systems demonstrating molecular self-assembly of two NDI (77, 78) species appended with tryptophan moieties, via π−π stacking interactions, hydrophobic interactions, and metal interactions. Both derivatives formed spherical aggregates from acetonitrile solution, while in 60% aqueous acetonitrile, (78) formed bundles of nanobelts, and (77) formed fractal-like assemblies (Figure 8(N,O).157 The group also conducted some important work on amino-aliphatic systems and also on the impact of solvent (solvophobic control). Specifically, they reported the hierarchical supramolecular self-assembly of Nsubstituted NDI using 2-aminooctane (84), by dissolving 84 in a good solvent (ethyl acetate), followed by fast dispersion into a poor solvent (methanol, methanol/water, or water), as shown in Figure 8(P−T).158 In their study, they observed that the use of a mixture of two poor solvents can provide f ine-tuning of intermolecular interactions. They observed a variety of nanostructures during this work, not only depending on solvent mixture, but also depending on concentration. At lower concentrations (0.5 mM) of (84) in a EtOAc/MeOH mixture, block-like structures of micrometer length and width ∼3.10 mM were produced, whereas at 1 mM, needle-like morphologies were produced. Interestingly, when the concentration was increased to 3 mM, flower-like structures with an average size of about 30 mm were observed. Knowing the role of solvent polarity in the self-assembly process, the group used water, and a water/ 11696

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 10. Substituted NDI core self-assembly with other aromatic systems resulting in supramolecular structures.

conditions. Interestingly, the end-capped species with elongated alkyl chain lengths in the NDI-dipeptide resulted in hydrogelation at lower pH compared to the corresponding shorter chains: NDI-dipeptide bearing end-capped propyl chains

formed hydrogels in neutral conditions, while relatively long n-octyl chains enhanced the ability to self-assemble in dilute solution. 11697

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

synthesized (79). These derivatives form gels in many organic (aromatic and aliphatic) solvents, but importantly, aggregationinduced enhanced emission (AIEE) effects are observed in aromatic hydrocarbon solvents, but not in the aliphatic gelstate.172 Thus, it was concluded that the formation of gels in organic hydrocarbons is due to four important features: (1) the NDI core, which promotes π−π stacking interactions; (2) the amide bonds of the amino acid, which provide hydrogen bonding functionality for self-association; (3) the aromatic ring of the phenylalanine residue, which facilitates π−π interactions, and (4) the long alkyl chains, which induce van der Waal’s interactions for gelation of the NDIs. Lin and co-workers synthesized four novel, asymmetric, NDIcapped hydrogelators, from a hydrophobic octyl-capped NDI dye, linked to dipeptides at the N-terminus (65), and investigated their self-assembling properties. Typically, the four dipeptides were prepared in conjugated form, such as NDI-Phe-Phe (1), NDI-Phe-Gly (2), NDI-Gly-Phe (3), and NDI-Gly-Gly (4), by solid-phase peptide synthesis (SPPS) using a 2-chlorotrityl chloride resin. Self-assembly of NDI, bearing Phe-Phe and Phe-Gly at the N-terminus, promoted the formation of 1D nanostructures and 3D colored hydrogels, both at acidic pH and at pH 7.4 (physiological conditions). They believe that the 1D nanostructured gels are stabilized through the intermolecular π−π interactions of the conjugated systems, and the extended hydrogen bonding of the dipeptide units. In addition, these water-soluble molecules were studied for cell imaging at concentrations below their minimum gel concentration.173 Further work relating to symmetry vs asymmetry, and aliphatics, was described by George and co-workers. They studied the dynamic self-assembly of NDI, bearing an oligoethylene chain on one side and an aliphatic chain at the other imide position (90), and which exhibited reversible vapochromism “turn-on” fluorescence, on exposure to CHCl3 vapor. Interestingly, microscopic analysis of the films, obtained by drying drop-cast solutions of 90, showed bundles of nanotapes in MeOH with CHCl3 (1%). It is believed that the reversible vapochromic effect originates from the self-assembly via J-aggregation.174 On the other hand, NDI bearing methoxytetraethylene-glycolamine (87) on both sides of the imide position self-assembled into organic nanoparticles from water/ methanol mixtures and exhibited a self-assembly-induced preassociated excimer (513 nm), i.e. enhanced green fluorescence.175 In another example, the group explored the effect of the linking group between the NDI imide and terminal functionality. Of note is their study on the role of the carbonate group in tuning molecular organization, and the resultant photophysical properties of the NDIs. In this regard, they synthesized two novel NDI derivatives bearing cholesteryl f unctional groups attached to the diimide position with a carbonate linkage (86) and an ether linkage (85), respectively. The NDI derivative bearing the carbonate-linkage (86) resulted in an unusual redshifted excimer emission (560 nm) in both the solid and the aggregated states, while the ether-linkage NDI derivative (85) showed typical NDI excimer emission (520 nm).176 It is also noted that NDI derivatives bearing cholesteryl through a carbonate-linkage form charge-transfer (CT) complexes with various electron-rich aromatic solvents, such as benzene, oxylene, and mesitylene, from an emissive ground state.177 Similar phenomena, demonstrating formation of one-dimensional supramolecular nanostructures through charge-transfer

A few important applications of these functionalized systems have been explored. First, Ramaiah and co-workers reported two new water-soluble amphiphilic conjugates based on naphthalene di- and monoimide chromophores (134, 135), and investigated their photophysical, self-assembly, and DNAbinding properties.168 These amphiphilic conjugates exhibited strong interactions with DNA and polynucleotides, for example poly(dG·dC)−poly(dG·dC) and poly(dA·dT)−poly(dA·dT), on the order of 5−8 × 10−4 M−1, which was evaluated by photo- and biophysical techniques. Interestingly, the amphiphilic conjugate (135) assembled into vesicular aggregates with an approximate diameter of 220 nm in aqueous medium, and these vesicular structures showed a strong affinity for hydrophobic molecules, such as Nile red. Furthermore, when exposed to DNA, the vesicles disassembled. Thus, formation of supramolecular assemblies and their encapsulation and release of hydrophobic dyes through DNA as a stimulus, and such methodology, can be used for real-world applications. Second, Müllen and co-workers reported two-dimensional supramolecular organic frameworks (2D SOF) through a host− guest complexation, in which NDI acted as an acceptor and Nmethyl viologenyl-substituted moieties as a donor (137), and consisting of a tris(methoxynaphthyl)-substituted truxene spacer (136) in combination with curcurbit[8]uril as host monomer toward monolayers (Figure 10). The UV−vis absorption, solid-state NMR, FT-IR, and wide-angle X-ray data confirmed successful complexation of all three monomers toward an internal long-range order with expected hexagonal superstructure. The obtained 2D SOF offers molecular selfassembly at a liquid−liquid interface, together with an exceptionally large-area, insoluble film monolayer with a thickness of 1.8 nm, and a homogeneous monolayer covering up to 0.25 cm2, along with free-standing overholes of 10 μm2, all of which were confirmed by TEM, AFM, and optical microscopy. This 2D SOF assembly opens up new doors for various applications, such as membranes, sensors, molecular sieves, and optoelectronic devices.169 3.3. Gelation and Solvophobic Controlled Assembly

In recent years, supramolecular gels, fabricated by noncovalent interactions, have attracted great interest, as organic gels are a very important form of soft matter with extensive potential applications. However, there are only very few examples of aromatic optically active gels obtained from planar gelators. The Banerjee group studied aggregation-induced enhancement of fluorescence and supramolecular gelation of NDI-appended peptides (80, 81) via solvophobic control (7:3, v/v of chloroform:MCH). Both compounds, (80) and (81), form gels in toluene, which on FE-SEM analysis were shown to be nanofibrillar in nature. For (80) the fibers ranged from 50 to 80 nm in width, while the fibers obtained from (81) were relatively thicker at 200 to 425 nm in width, and several micrometers in length.170 These derivatives formed gels in 4:1 v/v chloroform/ aromatic hydrocarbons, e.g. benzene, toluene, xylene, and mesitylene mixes, via CT complex formation.171 The role of different solvent molecules, with varying electron-donating capabilities, in the formation of CT complexes has been well-established through spectroscopic and computational studies. In particular, benzene, toluene, xylene (ortho, meta, and para), and mesitylene types of aromatic hydrocarbon have been used for the formation of CT complexes at room temperature. In another example, NDI bearing amino acid residues with long alkyl chains were 11698

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

(CT) assemblies of a coronene-based electron-donor (97), and of NDI derivative (98−100) as an electron-acceptor, have been observed. As expected, coronene−naphthalenediimide D−A pairs self-assemble into alternately stacked CT nanostructures with a high association constant.178 These results are important and illustrate how the choice of appropriate solvent(s) can assist development of organic electronic devices. In a similar vein, the Ghosh group synthesized simple NDI chromophores bearing carboxylic acid f unctionalities (93) for supramolecular assembly, through hydrogen-bonded J-aggregates, which exhibit white-light emission (0.33) with a very high fluorescence quantum yield (0.70) in the aggregated state in methylcyclohexane/CHCl3 (95:5, v/v) solvent mixtures.179 Their findings revealed that carboxylic acid groups play a crucial role in the mode of aggregation, the assembly, and the enhancement of luminescence (by 4 orders of magnitude),180 and showed that the assembly of the aggregates could be tuned from 2D sheets to 1D fibers by changing the number of carbon atoms (89, n = 0, 2, 3, 4) in between the imide and the carboxylic acid. They also synthesized a series of structurally related NDI chromophores appended with bis(trialkoxybenzamide) f unctionalities and studied their selfassembly and macroscopic properties. Here, the number of methylene groups between the NDI chromophore and the amide functionality were varied systematically. These derivatives selfassembled into unique nanostructured morphologies, including the following: short wires n = 0; nanowires n = 2: nanoribbons n = 3; and discontinuous nanof ibers n = 4, which were visualized by AFM, respectively. The tendency toward self-assembly was found to follow the following order: n = 0 ≫ n = 2 ≈ n = 3 > n = 4, which could be attributed to the varying H-bonding interactions. Furthermore, the morphological effects were highly relevant to the gelation properties of the individual chromophores. The most strongly aggregating n = 0 did not show gelation, while n = 2 and n = 3 showed the most promising gelation abilities in many common organic solvents and at very low concentrations ( 470 nm. Reproduced with permission from ref 303. Copyright 2013 the Royal Society of Chemistry.

Figure 46. Fluorescence imaging of (217) in skin melanoma cells. The image was acquired at the Cy5 light cube and transmitted mode with phase contrast objective after 40 min of incubation in a pH 5.2 phosphate buffer. Fluorescence imaging of (217) in skin melanoma cells. NucBlue Live ReadyProbes Reagent was added to the wells for the staining of the live cells. The images were acquired separately at the DAPI light cube, Cy5 light cube, and transmitted mode with phase contrast objective, after 60 min of incubation in a pH 4.5 phosphate buffer. Reproduced with permission from ref 304. Copyright 2016 the Royal Society of Chemistry.

Figure 47. (a) Family of UV−vis taken upon titration of (218)·4HCl (0.1 mM) with Cu(ClO4)2 at pH = 7.4 (T = 25 °C). Line with maximum absorption at 600 nm (violet line): initial spectrum, corresponding to 100% (218); Line tagged as λ1 (blue line): solution containing 76% [(218) Cu]+ species; Line tagged as λ2 (cyan line): final spectrum, taken upon addition of excess Cu(II) (10 equiv). Reproduced with permission from ref 305. Copyright 2015 Elsevier.

Their findings revealed the rate of the shuttling process of rotaxane (228) (Figure 53 (227−229)).324 The dynamics of the electrochemically induced shuttling in (228) were studied in detail using cyclic voltammetry (CV) at various scan rates and temperatures. The voltammogram of (228) in THF displays two reversible peaks associated with the consecutive reduction of NDI, as compared to the CV of the thread in (228). This reflects the stabilization of the mono- and dianionforms of NDI in the case of (228), which is due to the electrostatic nature of hydrogen bonding. Thus, both redox processes are shifted to less negative potentials, and the positive shift is notably larger in the case of the second reduction than the first (compare E1/2 = −0.64 and −0.98 V for 2a, and −0.68 and −1.21 V for (228)). Interestingly, the rate of shuttling for rotaxane (228) was 8 × 105 S−1, which is slightly slower than that in rotaxane (227) (1 × 106 S−1). The Stoddart group reported neutral NDI-based donor−acceptor [2]rotaxane (230), utilizing “click” chemistry. The solid-state structures of these derivatives were elucidated by X-ray crystallography. 1H NMR and electrochemical study revealed a redox-controllable “push-button” switching mechanism, i.e. that both geometrical

to produce novel redox-active mechanically interlocked structures. Herein, we will discuss recent examples of these types of structures. 5.1. Catenanes and Rotaxanes

In early 2008, the Leigh group reported electrochemically switchable [2]rotaxane-based molecular shuttles, which featured succinamide and NDI hydrogen-bonding stations on a benzylic-amide macrocycle, that can shuttle and switch between three different oxidation states, in solution and in a monolayer. Importantly, NDI exhibits a different binding affinity for the macrocycle at each oxidation state, and characterization of this redox-switching, which is occurring on the millisecond time scale, using cyclic voltammetry, is very convenient, as it can be performed with one experimental setup. Furthermore, the mild switching allows access to two of the states when the rotaxanes are confined to an alkanethiol-based self-assembled monolayer (SAM) on a gold electrode (Figure 53 (227, 228)).323 Furthermore, they have studied the formation of novel rotaxanes that incorporate PMI (229) and NDI (228) as redoxactive stations, and succinamide as a common template for ring formation and the preferred binding site for neutral molecules. 11724

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 48. Structural formulas and redox processes for (−)-NDI-Δ (219) and the reference NDI (220). Redox reactions between (−)-NDI-Δ, [(−)-NDI-Δ]3(•−)·3Li+, and [(−)-NDI-Δ]6−·6Li+. Redox reactions between NDI-ref, (220)•−·Li+, and (220)2−·2Li+. Each NDI unit is capable of undergoing two reversible one-electron redox processes. Molecular triangles composed of three NDI units undergo two reversible three-electron redox processes, amounting to a total of six electrons per molecule, whereas (220) can only accept two electrons. Reproduced with permission from ref 306. Copyright 2015 Wiley-VCH.

Figure 49. (a) Schematic illustration of the reaction of molecule (221) with hydrazine hydrate, and (b) confocal fluorescence microscopy images of the aggregates of (221) (a, b) before and (c, d) after adding hydrazine hydrate. Reproduced with permission from ref 311. Copyright 2015 the Royal Society of Chemistry.

Figure 50. “Turn-on” response of (222) toward Cu2+ ions. Reproduced with permission from ref 312. Copyright 2016 Elsevier Limited.

reorganization and redox-switching processes are occurring or can be induced in the [2]rotaxane.325 Figure 53 illustrates the shuttling mechanism of the [2]rotaxane based on the experimental support of cyclic voltammetry (CV) and scanning light scattering (SEC) analysis. It can be seen that, in the neutral form, the ring shuttles between NDI and one of the triazole units; however, the majority of its time is spent around the NDI unit. On the other hand, when NDI is reduced to NDI•−, the ring shuttles along the length of the dumbbell with little association with any site in particular. The group further demonstrated formation of mechanically interlocked dynamic oligorotaxanes 1D assemblies (232−235), in which the

Figure 51. Chemical structure of (223) compounds.

11725

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 52. (A) Nitrobenzene recognition by cyclophane bearing two NDI units linked by two alkyl chains (224). (B) Cartoon representations of the folding patterns targeted by compounds (225) and (226).

Figure 53. Chemical structures of molecular shuttles−thread complexes. Rotaxane and its dumbbell according to a convergent approach using the Huisgen 1,3-dipolar cycloaddition. Intermolecular Bingel reactions of [60]fullerene monoadducts bearing a NDI moiety with [60]fullerene successfully afforded novel [2]rotaxanes 1, while their intramolecular Bingel reactions gave [2]catenanes 2 with D−A−D−A stacking structure.

functional units communicate electronically through-space by way of π-orbital interactions. The oligorotaxanes reported

herein contain either one, two, three, or four NDI redox-active units and share electrons across the NDI stacks, which allows 11726

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 54. Structures and proton designations of host and guest molecules. Crystal structures of [2]pseudorotaxane: (A) PMDI2+ (237, n = 1)4− and (B) NDI2+ (237, n = 2)4−. Reproduced from ref 328. Copyright 2013 American Chemical Society.

Figure 55. Structure of [3]rotaxanes formed by (240, 241) and the mode of acid−base control of ICT between pyrene and NDI in the complex. Reproduced from 329. Copyright 2010 American Chemical Society.

the individual rotaxane components to move in and out of active electron-sharing domains and was evaluated by EPR and cyclic voltammetry (Figure 53).326 Kasai et al. reported fullerene-based [2]rotaxanes based on the donor−acceptor interaction between a fullerene derivative carrying an electrondeficient aromatic diimide moiety and an electron-donating dinaphthocrown ether.327 In the presence of DAN-38-crown-10 ether, the intermolecular Bingel reactions of fullerene (C60) monoadducts bearing a NDI moiety with fullerene successfully afforded novel [2]rotaxanes (231), while their intramolecular Bingel reactions gave [2]catenanes (236) with D−A−D−A

stacking structure. NMR, UV−vis, and mass spectroscopies were used to characterize the donor−acceptor [2]rotaxanes (Figure 53). Liu et al. studied the selective binding of tetrasulfonated 1,5-dinaphtho-38-crown-10 ((237, n = 2)4−) and tetrasulfonated 1,5-dinaphtho-32-crown-8 ((237, n = 1)4−) with PMI (238) and NDI (239) derivatives bearing cationic terminal groups (PMI2+ and NDI2+).328 In this study, they employed a new type of nonpyridinium guest, with watersoluble crown ethers used for the first time. These systems, based on water-soluble crown-ethers, may be useful for molecular devices in aqueous media (Figure 54). 11727

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 56. Proposed anion-induced displacement assay of interlocked structures incorporating NDI threads.330

Figure 57. Structures of host and guest components and the schematic representation of the probable orientation of (244) and (246) in [3]pseudorotaxane (247) in CH2Cl2.331

Figure 58. Movement process of [2]catenanes 5 under acid−base stimuli.

Furthermore, using click-chemistry, the group reported pHcontrolled intramolecular charge-transfer (ICT) in [3]rotaxanes compiled from electron-rich pyrene at the wheel (240), and an electron-deficient NDI unit in the middle of the axle (241).329 Interestingly, the present [3]rotaxanes display the ICT process controlled by the distance of the pyrene from the NDI moiety. This bistable [3]rotaxane is located at the ammonium of the aliphatic chain of the wheel at neutral pH, through a weaker CT band of pyrene and NDI, and at basic pH, the wheel moves from the secondary alkylanilinium to the triazolium, via efficient enhancement of the ICT intensity (Figure 55).

gives a colored pseudorotaxane assembly, and anion-complexation at the isophthalamide group of (243) causes displacement of the NDI and a resultant loss of color.330 The Das group described an NDI bearing two aza-crown ether (244) functionalities for the formation of [3]pseudorotaxanes (245) with imidazolium ions (246, 247) in nonpolar solvents, through H-bonding adduct formation. It was noted that, depending on the length of the linker to the imidazolium ion, the [3]pseudorotaxanes adopt different conformations with varying donor−acceptor interactions (Figure 57).331

5.2. Pseudorotaxanes

5.3. Catenanes

Beer et al. reported a supramolecular system for the formation of pseudorotaxnes of an electron-deficient NDI (242) and an electron-rich isophthalamide naphthohydroquinone (243) macrocycle, via sensing of anions through molecular motion (Figure 56). Interestingly, combination of (242) with (243)

Li et al. described a simple one-step method to synthesize [2]catenanes (248) bearing NDI (249) and crown ethers (250), by associated interaction templates. Importantly, Hbonding and π-donor−π-acceptor interactions led to the formation of this functional supramolecular system, in which 11728

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 59. Topological isomers that can form in a disulfide library generated from one single chiral LL building block (derived from L-cysteine) (251). The red components of the cartoons represent the NDI units, and the yellow dots represent the sulfur atoms connecting the building blocks. Reproduced from ref 333. Copyright 2014 American Chemical Society.

Figure 60. Schematic representation of the double plug−socket species assembley and disassembly under acid−base conditions.334

PET from anthracene to the NDI of the complexes (254) could be switched on and off by controlling reassembly, and the disassembly with an acid−base reaction, which is suggestive of a double plug−socket system (Figure 60).334

the reversible movement of the crown ether macrocycle (acid− base dependent) on the big ring between the two stations occurs via protonation and deprotonation (Figure 58).332 Very recently, the Sanders group described formation of various architectures such as knots, [2]catenane, and Solomonlinks, based on homochiral NDI (251) in water and utilizing a disulfide library of macrocycles containing topological isomers. Significantly, a mixture of racemic building blocks resulted in the near-quantitative formation of another new and more stable structure, assigned as a meso figure-eight knot (Figure 59).333 Liu et al. reported a novel “molecular key padlocks” concept, controlled through photooxidation, in which two robust divalent complexes were constructed using complementary anthracene vs NDI spacers (252), and two pairs of [24]crown8 ethers with secondary dialkylammonium functionalities (253) as binding motifs. Upon photoexcitation, the intermolecular

6. ION-CHANNELS Ion-channels are pores formed by membrane proteins whose functions include establishing electrical signals by gating the flow of ions across the cell membrane. Ion-channels are formed within the membranes of all cells and utilize two classes of proteins, such as ionophoric proteins and ion-transporters (sodium−potassium pump, sodium−calcium exchanger, etc.). To understand this natural phenomenon, Matile and coworkers described a very important strategy to synthesize organic rigid-rod molecules for the formation of synthetic ionchannels in lipid membranes. Rigid-rod-β-barrel-stave supra11729

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 61. Self-assembly of pore (A) from monomers (255) and a speculated suprastructure, consistent with experimental results and molecular dynamics simulations. (B) Molecular dynamics simulations of the pore-blocker complex of 243⊃244 in axial view with the bound blocker (256) in red. Reproduced with permission from ref 335. Copyright 2007 the Royal Society of Chemistry.

molecules (Figure 61(A)), composed of four p-octiphenyl staves, were brought together by β-sheet NDI hoops (255). The results from systematic screening of regioisomeric DAN amplifiers (256) demonstrate that the synthetic pores that operate by ion-pair-assisted π-clamping are capable of regioisomer recognition and respond to structural differences as subtle as the one between 2,6- and 2,7-DAN isomers (Figure 61).35,322,335 An alternative approach used self-assembly and self-organized rigid-rod π-helices, containing p-octiphenyls, with eight NDI units to form a closed channel while using ligand gating, i.e. intercalation with aromatic ligand (DAN) (Figure 61(B)), to untwist the helix-barrel to give open barrel stave ion-channels for anions in the lipid bilayer membrane. Matile and co-workers have also described a number of designs from monomeric to cyclic to rod-shaped NDI transporters (257), in which anion−π interactions take place between the anions and π-acidic aromatics.43 To achieve this, three NDIs were lined up in a rigid-rod oligomer, which allowed the anions to move along the π-acidic NDI surfaces within the lipid bilayer membranes (196) (Figure 33(B)).336 Results clearly show that π-acidification and active-site decrowding are found to govern the halide selectivity sequences, chloride > bromide > iodide, and supramolecular organization inverted acetate > nitrate to nitrate > acetate selectivity.44,337 Their results revealed that the anion−π interactions on monomeric surfaces are ideal for chloride recognition, whereas their supramolecular enhancement by π−π interactions appears perfect for nitrate. On the other hand, chloride transporters are relevant to treat channel opathies, and nitrate sensors to monitor cellular signaling and cardiovascular diseases. They have also introduced hydrophilic peptide anchoring to achieve high activity and high selectivity for anion−π-slides (Figure 62).43 Recently, the Matile group have used a layer-by-layer stacking concept to monitor photocurrent generation, and they found this to be greatly influenced by ions attached along the charge-transporting channel.338 Typically, the photocurrent generation was examined using the photosystem as the working electrode, a Pt wire as counter electrode, and an Ag/AgCl reference electrode, all immersed in a buffer containing a mobile hole-carrier, triethanolamine. As expected, the presence of a partially protonated system was found to be most significant for increased activity, implying that the system attracts holes and repels electrons, thus facilitating the photoinduced charge separation over long distances and hindering charge recombination at the same time, thus

Figure 62. Concept of anion−π slides. The “blue face” of the electrostatic potential surface of the O-NDI rod (257) supports the possibility of multi-ion hopping along the anion−π sites S1−S6 as indicated in the qualitative energy diagram.43,283 Reproduced from ref 283. Copyright 2006 American Chemical Society.

confirming the existence and significance of ion-gated photosystems. These findings may advance the use of organic electronics in organic field-effect transistors, light emitting diodes, and solar cells. The group also reported, for the first time, dynamic amphiphiles with fluorescent tails and covalent capture by hydrazide head groups based on NDI and PDI chromophores. Their findings confirmed that dynamic NDI (but not PDI due to poor partitioning) amphiphiles could be used to visualize cellular uptake pathways and to label intracellular membranes, and the further activity could be modulated by the composition of the mixed systems. Importantly, dynamic NDI amphiphiles partition selectively into the liquid-disordered (Ld) microdomains of mixed lipid bilayers and activate DNA as transporters.339 The Zhao group reported formation of trans-membrane nanopores in lipid membranes, made up of NDI (258) and pyrene-functionalized cyclic tricholate macrocycles (259) stacking over one another (Figure 63). Their findings revealed that, in the transmembrane pore formation of the oligo-cholate macrocycles, the acceptor−acceptor interacting NDI groups were more effective than the pyrene−NDI donor−acceptor interactions, and that such an arrangement allows the water molecules inside the macrocycles to interact with one another, to solvate the polar groups of the cholates, and also to exchange readily with the bulk water.340 11730

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 63. Aromatically functionalized cyclic tricholate macrocycles and aggregation to form a pore in a lipid bilayer membrane. Reproduced from ref 340. Copyright 2012 American Chemical Society.

Figure 64. Donor−acceptor triad of the photosynthetic reaction center.

core-substituted NDIs58−61 are rapidly emerging as powerful tools in creating functional nanomaterials, especially in biomimetic and bioinspired artificial systems, including electron- and energy-transfer systems,61 as well as in synthetic multifunctional pores.346 Recent developments in the synthesis of cNDIs have enabled several groups to probe the function of this interesting class of dye molecule in a molecular and supramolecular sense. One such approach is highlighted by Gust and Moore, in which the light-induced phenomena are mimicked by a molecular “triad” (Figure 58), consisting of an electron donor and acceptor linked to a photosensitive porphyrin group.347 This triad superstructure is incorporated into a lipid bilayer membrane, and upon excitation with light established a reduction potential near the outer surface of the bilayer and an oxidation potential near its inner surface mimicking natural phenomenon. We and others1,58,61,343 have reviewed this area; thus, we will only illustrate hereafter examples of the progress of NDIs and cNDIs in this field of study. To introduce this section, we have described a few early examples. Osuka et al. studied the photoinduced long-lived charge separation (0.14−80 μs) between a series of fixeddistance triads (13 and 17.2 Å), consisting of zinc porphyrin and NDI moieties, and bridged by aromatic spacers.348 In another example, the group synthesized a triad bearing a Zn(II)

7. ARTIFICIAL PHOTOSYNTHESIS Chemical systems capable of forming long-lived chargeseparated states are of great interest for advanced applications, including the following: conversion of solar energy into chemical potential; molecular-based optoelectronics; and the creation of artificial photosynthetic systems (Figure 64 (260, 261)).341 The bacterial photosynthetic reaction center (PRC) formed by long-lived light-induced charge separation shows several membrane proteins that fixate the functional components along an axis through a membrane within a 12 yoctolitre volume (12 × 10−24 L). The PRC has a number of attractive design features including the following: photoactivity, use of noncovalent bonding, and the function of charge separation, which plays a central role in molecular devices and catalytic developments.342 A significant part of this effort has been devoted to the study of photoinduced charge separation reactions, as a means of capturing and storing solar energy using covalent and noncovalent strategies, and involving donor−acceptor moieties.343 In recent years, NDIs have made a useful contribution to evaluation of these design principle, due to their ease of synthesis, their electron-donating and/or accepting properties,344,345 and their similarity to the naturally occurring acceptors in plant and bacterial photosynthetic reaction centers. Along with nonfluorescence NDIs,1 11731

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

porphyrin (ZnP) donor, with pyromellitimide (PMI) and NDI groups attached antipodally to the porphyrin, and studied electron transfer by control of the direction of intramolecular electron transfer between NDI and the pyromellitimide acceptors.349 In another example, the Wasielewski group reported photoinduced electron transfer, on a picosecond time-scale, within triad systems bearing NDIs as acceptor moieties.350,351 Furthermore, the use of NDIs in construction of artificial photosystems for solar energy conversion by using a noncovalent approach has also been evaluated. The Sessler group reported, for the first time, steady-state fluorescence quenching in a H-bonded rigid dyad, bearing porphyrin and NDI acceptor.352 Osuka et al. studied time-resolved picosecond transient absorption between NDI acceptor motifs directly attached to the meso position of Zn porphyrin (donor).353 Ghiggino et al. described formation of stable dyads of Zn porphyrin and pyridyl-NDI through coordination complexes; interestingly, photoexcitation of zinc porphyrin in nonpolar solvent resulted in charge separation with a lifetime of 1−100 μS.354 In the same year, Fukuzumi and co-workers reported the longest lived charge separation (450 μs) between ZnP-NDI covalently and noncovalently linked in solution. This is the first report of long-lived charge separation between donor−acceptor systems.355 Along with non-core-substituted NDIs, cNDIs also have shown breakthroughs in this field. In this regard, Chaignon et al. reported the synthesis and electron transfer in a Ru(bpy)3-cNDI-Ru(bpy)3 triad system.356 Earlier, Matile and co-workers described the first breakthrough in cNDI chemistry through the creation of an artificial photosynthetic system: a tetrameric M-helix based on cNDIs that is long enough to span a lipid bilayer membrane and undergo symmetry-breaking ultrafast and quantitative photoinduced charge separation.62 They also studied ultrafast photoinduced charge separation in NDI-based systems in liquid solution and in a lipid membrane. The lifetime of the charge-separated state was found to increase from 22 to 45 ps by going from bi- to octachromophoric blue systems in methanol, while a 400 ps decay component was observed in the lipid membrane. This lifetime lengthening is explained in terms of charge migration, and is most efficient when the octachromophoric systems are assembled as supramolecular tetramers in the lipid membrane, and the average charge-separated state lifetime of the red system in methanol is even larger, at about 750 ps.357,358 The details of this study have been reviewed by the Matile group359 and also discussed in detail in our earlier review.57 Herein, we provide a brief update on the novel structures produced and their photophysical studies. The zipper and layer-by-layer assembly of colorful NDIs utilized for artificial photosystems (183, 184) are illustrated as reported by Matile and co-workers (Figure 65).359−361 The effect of substituents on the optical properties of NDI derivatives was also recently studied by the Bickelhaupt group using frontier orbital analysis.362 Charge transfer is also affected by the molecular conformation and intramolecular π-stacking due the D−A interaction. Sao et al.364 investigated a unique intramolecular D−A interaction in a pair of regioisomeric covalently bridged NDI dimers. Though the D−A interaction of the ortho positioned NDI dimer is characterized by a weak absorbance, it exclusively dominates the photoluminescence (PL) properties of the dimers. Vibronically resolved PL and PL excitation spectra exhibiting mirror-image symmetry, a negligible Stokes shift, and a weak solvatochromism point to a partial chargetransfer character of the D−A state. Using F− binding, they

Figure 65. Building blocks for oriented multicolored antiparallel redox gradient−supramolecular n/p-heterojunction (OMARG-SHJ) photosystems. Frontier orbital energy levels of NDIs (C2-G), p-oligophenyls (POPs), and oligophenylethynyls (OPEs) (solid lines, HOMO; dashed lines, LUMO; dashed arrows, absorption of light; with wavelength (nm) of maximal absorption (top) and emission (bottom)). Reproduced with permission from ref 258. Copyright 2010 Wiley-VCH. Reproduced with permission from ref 363. Copyright 2009 the Royal Chemical Society.

established that the D−A interaction involves one donor (xylylene ring) and a pair of noninteracting acceptor NDI units, where no charge transfer can be observed (see Figure 66).

Figure 66. Schematic summarizes the (A) evolution of PL as an acceptor−donor−acceptor (A−D−A) triad system folds into an intramolecular H-type aggregate, and (B) the competition between F− and the aromatic ring donors. Reproduced with permission from ref 364. Copyright 2016 the Royal Chemical Society.

Majima’s group studied the intramolecular electron transfer (ET) in PDI dimer and PDI−pyromellitimide (PI) and PDI− NDI dyads using femtosecond laser flash photolysis. They found that efficient intramolecular ET occurred in PDI−PI and PDI−NDI because of the sufficient driving forces. PDI•−*− PDI and PDI•−*−PDI•− exhibited different ET pathways.365 In a typical zipper assembly, cNDI-chromophores attached to POP or OPE chromophores (Figure 28 (183, 184)) and are assembled step-by-step to build mutually interdigitating πstacks along interdigitating rigid-rod scaffolds, as shown in Figure 67. Furthermore, the group have made significant contributions by building multilayer zipper and LbL assemblies for artificial photosystems based on multicolor cNDIs, attached to an oligophenylethynyl (OPE-NDI) or p-oligophenyl (POPNDI) backbone, in the dry and wet states on solid substrates, 11732

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 67. Zipper assembly of notional SHJ architectures; see Figure 28 (183, 184).358 Reproduced from ref 358. Copyright 2009 American Chemical Society.

from the excited porphyrin to the cNDI was found to take place with ∼1 ps and ∼25 ps time constants in the ZnP and FbP triads, respectively, along with CR with a time constant of about ∼7 and 16 ps. The buildup of the CS state population in the ZnP triad is independent of the excitation wavelength, indicating that charge separation takes place from the lowest singlet excited state. The Pascu group reported new donor− acceptor hybrids of Zn(II)-metalated 5,15-diaryl porphyrins with the electron−acceptor molecule di-n-hexyl N-substituted 1,2,4,8-NDI. Variation of the porphyrin−NDI binding strengths was determined in the solid-state by synchrotron Xray diffraction and probed in solution using 1H NMR spectroscopy.370 It was found that the 5,15-diaryl substituent of the symmetrical porphyrins progressively enhanced the binding affinity of π-complexation with a dialkyl N-substituted NDI. They concluded that the presence of electron-donating or -withdrawing substituents on the meso-phenyls peripheral to the highly conjugated nature of the porphyrin ring can affect the physical and chemical properties of the macrocycle, and hence the strength of the interactions within the donor− acceptor complex in solution, and the nature of the aromatic stacking in the resulting complex in the solid-state. Very recently, we have also synthesized multichromophoric arrays, i.e. pentads (265), composed of a cNDI as a center core and four ZnP or FbP substituted on the naphthalene core via aniline bridges, and studied their electron transfer between porphyrins and NDIs using a combination of stationary and ultrafast spectroscopy (Figure 68). Results reveal that, in a polar solvent, charge recombination in the charge-separated state and the ground state occurs within a few picoseconds.371 In this study, we used the terminal S1 state, for the state populated upon excitation in the lowest energy band, whereas those populated upon excitation in the Q and B bands were named local Q and B states (LQ and LB), respectively. Indeed, the fluorescing excited state is delocalized over the whole array of pentads, whereas the other upper states are localized on a single porphyrin (i.e., the LQ states) or partially delocalized over pairs of adjacent porphyrins (i.e., the LB states). Therefore, these pentads can be considered as near-IR cNDI dyes. To-date, these are the smallest S1−S0 gaps reported for c-NDIs (Figure 70(A)). We have also synthesized and investigated the excitedstate dynamics of two porphyrin−naphthalenediimide dyads connected via a peptide linker to a core-substituted NDI (266) using optical spectroscopy (Figure 69). These dyads exhibit rich photophysics due to the large number of electronic excited states below 3 eV. In the case of dyad (266, M = Zn) in apolar

successively using a quartz crystal microbalance (QCM) and surface plasmon resonance (SPR).363 Recently, we reported the synthesis of rigid and planar structures (262−265) to study photoinduced electron transfer, based on porphyrin and tetrathiafulvalene (TTF) as a donor and cNDI as an acceptor, functionalized via a diaza rigid bridge, amine, or C−C bond formation (Figure 68). Further electrochemical and photophysical testing demonstrated the combination of pronounced redox behavior and a strong photoinduced intramolecular charge transfer within the NDITTF (262) dyad. Typically, a significant π−π interaction, with near IR absorption (1270 nm), along with a 852 nm (cNDITTF)•+ band, was observed upon chemical oxidation with FeCl3.366 In another example, we described femtosecondresolved spectroscopy of the two dyads consisting of free-base or zinc-tetraphenylporphyrin (FbTPP, ZnTPP) attached to a tetra-cNDI unit via a 2,3-annulated linkage (263). 367 Interestingly, the reported dyads are very strongly coupled and behave like new chemical entities with distinct stationary absorption and emission spectra and excited-state dynamics. Upon photoexcitation of the dyad in dichloromethane, the charge separation (CS) from the ZnTPP to the cNDI unit was in the range 1−3 ps; however, charge recombination (CR) occurred within 10−20 ps depending on the solvent used. Ogawa and co-workers described the synthesis of four types of donor (porphyrin)−acceptor (NDI) dyads linked through an imide linkage. Typically, porphyrins bear different central metals (zinc or rhodium) with different substituents on the other side of porphyrin macrocycles (tert-butyl or methoxy), and their capacity to act as a single molecular diode is studied. Their findings proved that the effect of the central metal was very significant, as the zinc dyad showed a high level of CT character, whereas that of the rhodium dyads indicated insignificant charge-transfer character.368 In another report, we have studied the excited-state dynamics, i.e. ultrafast fluorescence and transient absorption spectroscopy, of two triads (FbTPP-NDI-FbP and ZnTPP-cNDI-ZnTPP (264)).369 The absorption spectra of the triads are almost the composite of those of the constituents, pointing to a weak electronic coupling and to a localization of the excitation energy on one of the porphyrins. In cyclohexane, the excited-state dynamics of the triads are essentially the same as those of the individual tetraphenylporphyrins, with the exception of the Soret emission of the ZnP triad, whose lifetime exhibits a more than 10-fold shortening compared to Zn-tetraphenylporphyrin. However, in tetrahydrofuran and benzonitrile solvents, charge separation 11733

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 68. Charge separation at the HOMO → LUMO transition where the NDI core acts as acceptor and other moieties as donors. Reproduced with permission from ref 366. Copyright 2011 Elsevier. Reproduced with permission from ref 367, 369, and 371. Copyright 2011, 2013, and 2014 PCCP Owner Societies.

solvents, excitation energy transfer from the vibrationally hot singlet excited porphyrin (free-base) to the NDI takes place with a 500 fs time constant. Importantly, electronic energy ends up in the NDI-localized triplet state, which decays to the ground state on a microsecond time scale. However, in polar solvents, ground-state recovery is faster by 5 orders of magnitude because of the occurrence of charge separation followed by recombination. On the other hand, much less solvent dependence is observed with dyad (266, M = 2H), where charge separation is endergonic and occurs on the same

time scale as intersystem crossing. A very interesting feature of the dyad is that excitation energy transfer takes place in the opposite direction, i.e. from the NDI to the porphyrin, which then undergoes intersystem crossing to the triplet state, followed by triplet energy transfer back to the NDI (Figure 69).372 In a similar direction, the Ghiggino group reported a photoinduced electron transfer in a covalently linked donor− acceptor system, i.e. zinc-porphyrin (donor) and cNDI (acceptor). Typically, the fluorescence of ZnTPP-cNDI is 11734

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 69. Chemical structure of the photoexcited porphyrin−naphthalenediimide dyad, and its excited-state dynamics. Reproduced from ref 372. Copyright 2015 American Chemical Society.

Figure 70. (A) Energy-level scheme of (265) (the possible splitting of the LB level due to decreased symmetry and excitonic interaction has been neglected). (B) Kinetic scheme employed in the Global Target Analysis of the transient absorption data of the ZnTPP-amino-substituted NDI dyad. Reproduced from ref 371 with permission from the PCCP Owner Societies. Reproduced with permission from ref 373. Copyright 2013 Elsevier.

be oxidized with PbO2 in dichloromethane via fluorescence quenching, and further fluorescence emission can be restored on the addition of saturated NaBH 4 solution. This interconversion can be repeated several times and may make them an excellent candidate for redox switches (Figure 72).375 The Nocera group studied electron transfer in noncovalently assembled donor and acceptor systems, i.e. ZnP (273) and NDI via two-point H-bonding between the amidinium of the porphyrin and the carboxylate (274) or sulfonate (275) functionalities of NDIs (Figure 73).376 Their findings revealed that proton-coupled electron transfer between H-bonded D−A retards the ET rate by over 2 orders of magnitude. Work by Odobel and Hammarstrom reported dyads and triades bearing ZnP as a donor and NDI and NI units as electron acceptors (276−279).377 They demonstrated the state-selective direction of electron transfer, when they observed an excited S1 state of the porphyrin Q-bands by electron transfer to the NDI unit, while excitation of the S2 state, i.e. the Soret band of the porphyrin, occurred by electron transfer to the NI moieties. Favereau et al.378 designed a molecular tetrad system (280) to mimic the natural oxygenic photosynthesis “Z-scheme” system. This tetrad is composed of two dyes 4,4-difluoro1,3,5,7-tetramethyl-2,6-diethyl-4-bora-3a,4a-diaza-s-indacene (Bodipy) and a RuII(bipyridine)3 which are connected to an NDI electron acceptor and tetraalkylphenyldiamine (TAPD) playing the role of electron donor. The Z-scheme is a unique system able to generate both a strong oxidant and a strong reductant using two photons of the visible spectrum via a single charge-separated state (see Figure 74). Fukuzumi and Kay described the photoinduced electron transfer between a silicon phthalocyanine (donor) with two units of acceptor NDI, with (281) and without fullerene (C60),

quenched in both polar and nonpolar solvents, i.e. benzonitrile and toluene, respectively (Figure 70(B)).373 Ultrafast transient absorption studies have attributed these observations to the formation of the ZnTPP•+ radical cation, and the cNDI•− radical anion. They have also shown that electron transfer can occur directly upon excitation to the S2 state of the ZnTPP, which is followed by rapid charge recombination to form S1 and subsequently a further slower ET to cNDI. The kinetics of CS and charge CR are strongly solvent dependent, as in the more polar solvent benzonitrile, the accelerated charge recombination rate (kCR) (kCR = 1.59 × 1011 s−1) was faster than in toluene (kCR = 8 × 109 s−1), in which intersystem crossing (ISC) from the CS state to form the lowest energy porphyrin triplet state was the dominant decay pathway (kCR/ISC = 3.46 × 1010 s−1). The consequences of these results for designing molecular systems to potentially utilize ET following S2 excitation are discussed. Wasielewski and co-workers reported a series of PDIs and NDIs bearing varying degrees of trifluoromethylation to the core, and they studied their spectroelectrochemical and EPR spectroscopy. UV−vis absorption and fluorescence spectroscopy were used to determine their radical anions. Their studies revealed that trifluoromethylation of the imides and diimides caused their one-electron reduction potentials to be substantially more positive, relative to those of their unsubstituted counterparts (Figure 71).374 Maniam et al. reported the synthesis and spectroscopic properties of a push−pull cNDI bearing o-phenylenediamine on one side, and cyanoethene dithiolate on the other side (271). Their findings revealed that the push−pull NDI underwent reversible molecular switch action via oxidation and reduction processes. Interestingly, this derivative can also 11735

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

fullerene or NDI as electron acceptors to the TPA-BDP-ZnP triad through zinc−nitrogen axial coordination resulted in electron donor−acceptor polyads. Photoexcitation of ZnP resulted in rapid electron transfer with a rate of 5.0 × 109 − 3.5 × 1010 s−1 to the coordinated C60 or NDI, yielding a CS ion-pair species in nonpolar solvent (toluene). Crucially, in the presence of C60, or NDI, charge-separated states were observed and importantly fast-forward ET and relatively slow back ET occurred in these new systems (Figure 76). The Zhao group synthesized two NDI-C60 dyads: one with phenyl (286) and another with an ethylene glycol spacer (287), that showed strong absorption of visible light.381 The photophysical properties of both dyads were studied with steady-state and time-resolved spectroscopy. Alkylaminosubstituted cNDI was used as a light-harvesting antenna, and the C60 unit was used as a spin converter for the ISC from singlet state to triplet excited state. In their study, on visible light photoexcitation, the triplet excited states of the dyads were populated with a lifetime up to 90.1 μs. These types of dyads can be used as general structural motifs of heavy-atom (Br, I)free organic triplet photosensitizers, for triplet−triplet annihilation (TTA) based up-conversion, in which the photophysical properties can be easily changed, using different light-harvesting antennae, and the ISC properties readily predicted (Figure 77). The effect of the light-harvesting and the lifetimes of the sensitizers on the efficiency of TTA upconversion are shown in Figure 77 (the vibration energy levels of each electronic state are omitted for clarity. Poddutoori et al. reported two self-assembled donor− acceptor triads (288−290) consisting of aluminum porphyrin (AlP), as a donor, and axially bound NDI as an acceptor, along with tetrathiafulvalene (TTF) as a secondary donor, as shown in Figure 78. Importantly, the NDI and TTF units are attached to AlP on the opposite side through a covalent and a coordination bond, respectively. The results showed that fluorescence quenching of the lowest excited singlet state of the porphyrin, through electron transfer to NDI and holetransfer to TTF, depends on the solvents used, as in dichloromethane hole transfer to TTF dominates, but in benzonitrile electron transfer to NDI occurred. The photoinduced electron transfer within the triad in benzonitrile gives a singlet state lifetime of TTF•+−NDI•− within 200−300 ns, and a triplet state lifetime ∼10 μS; however, in dicholoromethane the lifetime of the charge separation is 740 nm), which would enable them to be used as building blocks for molecular applications, such as dual photon imaging, switch on−off, or even photodynamic therapy.497 We have studied a similar system of a donor−π-bridge−acceptor (D−π−A) module where a simple triphenylamine functionality served as an

Figure 123. Chemical structures of the arylenediimide-thiophene derivatives.491 11766

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

power conversion efficiency, and the electron lifetime of the diad in comparison with the analogue where the NDI moiety is absent in a dye-sensitized solar cell.498 In another report, Fernando et al. investigated the optical and electrochemical properties of the styryl-substitution on NDI’s (441, 442) through the imide linkage, using 2-ethylhexylamino and 5-(2-ethylhexyl)thiophene as substitution analogues on the core as well as 2-ethyl-hexyl, 4-thienylphenyl, and hydrogen as substituents to the imides. Interestingly, they found that the overall optoelectronic properties of the NDI derivative decreased as the donating effects of the core substituent were increased. They further confirmed that the imide substituent also plays a role in the molecular packing, hence film properties and bulk hetero junctions vary. Their findings clearly show that, when the core is substituted by a styryl imide, this leads to greater absorption spectrum, yet again adding to the arsenal of design motifs that can be used in conjunction with the everversatile NDI.499 As has been discussed for organic-electronic devices, low lying LUMO levels are generally beneficial. Thus, Kumar et al. synthesized extraordinarily electron-deficient NDIs (443) bearing triphenyl phosphonium on the core, which form very stable radicals.500 These NDI radicles, the first of their kind, are incredibly stable in air and are unphased by silica gel, which was used in the process of purification. Their electrontransfer properties may also lead to interesting uses as initiators in various chemical reactions, as the stability of the radical would mean an “off the shelf” reagent would be available to

Figure 125. Optical microscope images of (A) orange needles ((431)· 2 MeOH) and (B) red-purple needles (431) and (C) the color changes of (431) after exposure to saturated vapors of various volatile organic compounds. Reproduced with permission from ref 494. Copyright 2010 Wiley-VCH.

electron donor, a cyanoacrylic acid as an electron acceptor and anchoring group, and a NDI moiety as the π-bridge embedded between the two thiophene units. We showed that the incorporation of the NDI moiety as a strong conjugating functionality improved light-harvesting, photocurrent density,

Figure 126. Donor-acceptor systems based on 2,6-core-substituted NDI derivatives. 11767

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

chemists. The NDI radical is of significant interest due to its applicability in a myriad of fields, not only due to the phosphonium group and associated spectator ion but also due to the potential to enhance the stability of electron-deficient compounds. This is a common problem in the design and synthesis of organic-electronic materials. These properties would be heavily suited toward applications as switchable panchromatic materials, as well as for donor−acceptor systems used within organic solar cells, OTFTs as well as OLEDs (Figure 127).

junctions. Aromatic units were used at the terminal ends of the OPE in order to “anchor” itself to the CNT via London forces. The optical properties of these compounds were studied by drop casting the compound onto graphene.76 Wu et al. reported the synthesis and characterization of an interesting rigid cofacial NDI dimer (451).83 The NDI units were separated by a distance of 3.5 Å from one another, which is less than the usual 4.5 Å that is seen in standard mono-NDI compounds (Figure 128). The closeness of the two NDI units

Figure 127. Phosphonium-substituted NDI radical cation.

Figure 128. Tubular representations of the rigid cyclophane dimer (−)-2DNI plan, and side-on views of solid-state structures of (−)-2NDI, showing the close intramolecular distance and twist angle between the two. Reproduced with permission from ref 83. Copyright 2014 the Wiley-VCH.

Li et al. synthesized a series of core-substituted NDI derivatives with arylacetylenes and demonstrated their dual functions because of their n-type semiconducting and waveguiding properties. The derivatives (444−447) show waveguiding properties when assembled into rods and importantly show an aggregate-induced emission (AIE) effect with quantum efficiencies reaching almost 10%. They also found that the emission of the NDI derivatives could also be tuned by changing the aryl group. All compounds displayed some level of OFET behavior; however, only compounds (444−446) with C20H41 alkyl chain substitution were significantly stability in air, while (444−446) with C6H13 alkyl chain substitution showed n-type properties under inert atmospheric conditions. The electron mobilities of the compounds, in film, were enhanced with thermal annealing, and the highest recorded electron mobility was for compound (446): with C6H13 as the alkyl, electron mobility was 0.075 cm2/V·s (Figure 126).501 Polander et al. explored the affects the 2,6-substitution on NDI has on the overall electrochemical and optical properties. The synthetic procedures in this paper are also worth noting, as the intermediate compounds, 2,6-vinyl NDI, 2,6-formyl NDI, and N,N′-di(n-hexyl)-2,6-bis(2,2-dicyanovinyl) NDI (448) are of synthetic significance, given their use in a variety of chemical reactions.502 Most modification to NDI exists around the substitution of the amide as well as to the aromatic core; this is done due to the ease and versatility it offers. However, Etheridge et al. reported synthesis of thionated-NDI (449) through imide oxygen and substitution effects on the optical and electrical properties of (449). It, (449), showed a dramatic red-shift in UV−vis absorption and consequently much lower optical bandgap. Interestingly, the bathochromic shift increases with ∼50 nm per sulfur increase to the imide oxygen. This study clearly demonstrated that, along with core-substitution, imide-thionation of carbonyls is an effectively simple way to tune NDI for use in optoelectronic applications in future prospectives.503 Grunder et al. reported the use of NDI (450) embedded within an oligo(phenylene ethylene system) with a varying number of subunits in order to create “molecular wires” of different lengths, from 2.3 to 6.6 nm, in order to study their ability to bridge a gap between carbon nanotubes acting as

led to efficient π overlap and showed very interesting photophysical and electrochemical properties. The UV−vis absorption spectroscopy of (451) demonstrated that the dimer is strongly exciton-coupled, while fluorescence spectroscopy demonstrated a long-lived charge separation state (20 ns) along with a large Stokes shift and a lack of vibronic features. They have also shown that when excited from the π−π* molecular orbital, it is shared between both the NDI units, leading to an emissive excimer-like state, indicating electron delocalization in the π* orbitals. CV experiments clearly showed the dimer has four distinct redox states, each of which is isolatable in boththe solution and the solid phase. These properties are key to the development of better materials for use in organic electronic applications. Gabutti et al. also reported on the synthesis of cofacial NDI dimer separated by meta-dimethylenebenzene spacers (452), as shown in Figure 129.504 These types of dimers, given their proximity, exhibit Förster resonance energy transfer (FRET). The FRET occurs through dipole−dipole interactions and is therefore very sensitive to changes in distance, as well as the optical characteristics, such as spectral overlap of the donor and acceptor units. Having one chromophore with the ability to be “tuned” via core-substitution is of great benefit in the study of such phenomena. They found that intramolecular FRET occurred for the methoxy-substituted NDI derivative, but was inhibited by the sulfanyl and piperidinyl substitution. The synthesis of potent nonfullerene acceptors is possibly one of the greatest challenges in developing new materials for use in organic photovoltaic devices. In the area of small molecules that perform this task, NDI has proven to be very useful. Liu et al. synthesized a simple, yet effective, NDI dimer (453) linked simply through a vinyl linkage and evaluated its applicability for OTFT. Impressively, (453) gave an electron mobility of 0.0335 cm2/V·s, under ambient conditions in air as spun; upon annealing at 120 °C, this increased to 0.365 cm2/V·s. The OPV device was prepared in a 1:1 ratio with PTB7 as the donor and 11768

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 129. NDI oligomers and NDIs imbedded in copolymer oligomers.

solar cells (Figure 130; compounds are labeled NDI-nTH (n = 1, 2, 3, 4) and NDI-nT (n = 2, 3)). The optical bandgap for these compounds varied from 1.4 to 2.1 eV, which would suggest an affinity for light harvesting as well as exciton generation. IPCE data showed a broad range of absorption with a maximum efficiency at 16%. From OFET devices made from the NDI polymers electron mobility varied from 5 × 10−6 to 9 × 10−4 for electron mobility and 2.2 × 10−5 to 2.1 × 10−3. In a blended film with P3HT and NDI derivative, in a 1:3 mixture ratio respectively and efficiency of 1.5% was recorded with an open gate voltage of 0.82 V.506 Hwang et al. synthesized a novel NDI dimer (456) linked via a bis(thienyl)tetrazine bridge and prepared several OFET devices by spin-casting, as well as inkjet printing. From data gathered from film investigation, dichlorobenzene was chosen for optimal morphology. The spin-cast devices gave an electron mobility of 0.15 cm2/V·s, whereas the inkjet printed devices, on flexible plastic substrates, gave 0.17 cm2/V·s. The devices showed considerable resilience to multiple scans and showed little degradation under operation under ambient conditions as

the NDI dimer as the acceptor with a doping of 0.5% 1,8diiodooctane (DIO). The highest recorded efficiency was found to be 2.41%, the highest, at the time, for any known NDI small nonfullerene acceptor.505 Interesting platinum containing acetylide oligomers with varying length, capped with NDI (454), were synthesized using TIPS and HOM synthesis techniques by Keller et al. The compounds displayed both electro- and photoactivity and were characterized using CV, UV−vis absorption, and fluorescence spectroscopy. From the CV and UV data, they confirmed the presence of a triplet exciton, in which the formation of an oxidized platinum chain and the reduction of one of the NDI acceptors were shown. However, fluorescence studies showed the triplet excited state was quenchable via a nonradiative pathway prior to phosphorescence occurring. This study gave valuable insight into the relationship between organometallic structures and the mechanisms of charge transport that exist in chromophores.233 Ahmed et al. synthesized a polymeric NDI (455) based electron acceptor with varying length of oligothiophene πspacer (number of thiophenes (1, 2, 3, 4)) to use in organic 11769

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

polymers. The measured electron mobility values were 0.06 and 0.002 cm 2 /V·s for the poly-NDI and poly-PDI, respectively. It was concluded that although the electron mobilities are moderate, the poly-NDI demonstrated the possibility of stable n-channel polymers, and despite their large electron deficiency, they remained stable at ambient conditions in air and were still functional after 12 weeks. This opened many avenues toward the use of NDI-based potent polymeric materials for use in OTFT devices. Kim and coworkers synthesized a series of n-type NDI-thiophene copolymer acceptors by Stille coupling of 4,9-dibromo-2,7bis(alkyl)benzo[lmn][3,8]-phenanthroline-1,3,6,8-tetraone and 2,5-bis(trimethylstannyl)thiophene, with all of the polymers bearing different side chains at the imide-position, i.e., 2hexyldecyl (HD), 2-octyldodecyl (OD), and 2-decyltetradecyl (DT) groups (459a; m = 1 and alkyl N-substituted 460, 461). Furthermore, the photovoltaic characteristics of the optimal BHJ morphology generate highly efficient all-PSCs with PCEs approaching 6%.511,512 In another report, Xia et al. used an NDI (460) acceptor polymer in combination with quinoxaline− thiophene as a donor polymer for fabrication of inverted allpolymer solar cells devices.513 Kurosawa et al. reported the impact polystyrene side chains have on NDI (460) n-type polymer semiconductors in field-effect transistors.514 They observed that all the polymers showed similar high electron mobilities (0.2 cm2/V·s). An et al.515 conducted a comparative performance study of (459b and 462) polymers with hybrid siloxane pentyl chains as N-substitution in OFET devices. They found that the solvents from which polymers were spin-coated to fabricate the OFET devices had a significant effect on device performance due to polymer packing and film morphologies. The use of chloroform has improved the electron transport efficiencies of (459b and 462) to 0.21 ± 0.05 and 0.16 ± 0.01 cm2/ V·s, respectively: 1 order of magnitude higher in comparison to samples prepared from other solvents. The Zhou group described a comparative study of two solutionprocessable polymer solar cells (PSCs), using NDI (PCPDTNDI) (463) or PDI (PCPDT−PDI) (487) polymers as the electron acceptor and PTB7 as the donor polymer. In comparison, PCPDT−PDI exhibited better photovoltaic performance with a PCE of 2.13% when using 1-chloronaphthalene (CN) as an additive to obtain a good film morphology and to improve the electron mobility.516 However, NDIdithienosilole (NDI-PDTS) alternate copolymer (464) gave a PCE of 1.19% by adding 1,8-diiodooctane (DIO) as a solvent additive.517 Wu et al.518 reported the synthesis and solar cell device fabrication of water-soluble single-junction NDI based polymers (465−467) for the first time. The water solubility of these derivatives came from the conjugated main chain, with highly polar amino- or ammonium-functionalized side chains, respectively, and provides them with good processability in alcohol. Both the NDIs and amino- (465) and bromidequaternized ammonium groups (466) allow for work function tuning effects at the interface of the metal electrode and ndoping effects at the interface with the active layer of the acceptor. Interestingly, derivative (465) also exhibited photoinduced conductivity; on the other hand, (466) self-doped without excitation. Because of their π-delocalized planar structure and electron transporting capabilities, the resulting devices exhibit promising performance with power conversion efficiencies exceeding the 10% performance milestone. The specific properties of these kinds of water/alcohol soluble derivatives make them promising candidates for efficient

Figure 130. Schematic of a typical donor/acceptor interface in a BHJ solar cell with associated HOMO/LUMO energy level offsets, and below the HOMO/LUMO energy levels of oligothiophene-functionalized NDIs compared to those of poly(3-hexylthiophene) (P3HT). Reproduced from ref 506. Copyright 2011 American Chemical Society.

well as in vacuum, annealing rectified positive shifts in threshold voltages. This study clearly demonstrated that simple NDI dimers could be incorporated into inkjet-printing, a real step forward toward efficient and cheap organic electronic devices.507 In another report by Polander et al., a simple and efficient synthetic strategy was used for the preparation of a variety of NDI (457, 458) derivatives.508 CV results clearly demonstrated the formation of either 4 or 6 distinct reversible redox peaks from reduction of each NDI to its anion followed by further reduction to the dianion. The di-NDI was shown to have an electron mobility of 0.34 cm2/V·s, and the tri-NDI displayed 0.014 cm2/V·s. Although the mobility values are modest, the synthetic strategies are worth noting considering their ease and broad application in the synthesis of NDI containing organo-electronic materials. Takai et al.509 studied a ferrocene (Fc) N-substituted NDI, where ferrocenes act as electron donors to the electrondeficient NDI. The Fc units provide high crystallinity to FcNDI-Fc with good solubility in conventional organic solvents. This compound exhibited broad charge transfer absorption from the UV region to 1500 nm in the solid state. 9.2. NDI Based Polymers

Polymer solar cells are growing in interest due to their high efficiency and stability; however, they suffer from tedious synthetic difficulties. To evaluate the possibility of using NDIbased polymers in solar cells, Chen et al. synthesized D−A polymers based on NDI (459b, m = 2) and PDI (486).510 The NDI and PDI were chosen as a result of their large electron affinity and as a means to investigate the structural activity relationship that exists. The compounds were synthesized starting from their corresponding bistannyl thiophene via Stille coupling. Each diimide unit was separated by a twined thiophene π-spacer (Figure 131). OTFTs were constructed using standard bottom-gate architecture from both the 11770

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 131. Polymerized NDIs, and the role of the spacer on optoelectronic properties of NDI polymers. 11771

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

absorption of the polymer was seen across the visible spectrum and toward the near IR. CV data suggested the reversible formation of stable, anion and subsequent dianion species. The optical and electrochemical properties of this simple NDI polymer, as well as the ease of synthesis, could see future use of NDI polymers in organic electronics. Kim et al.527 synthesized a series of n-type D−A terpolymers composed of electron-deficient NDI-based units conjugated to two electron-rich thiophene and selenophene units to control the crystalline behavior of NDI-based polymer acceptors and enhance the performance of all-polymer solar cells. They have systematically controlled the crystallinity of the polymer by varying the thiophene/selenophene ratio in the polymer backbone and found enhanced terpolymer crystallinity with increasing selenophene content. This has led to significant enhancement of electron mobility and a significant increase in the power conversion efficiency values from 2.50 at 1:0 thiophene/selenophene ratio to 3.60% at 0:1 thiophene/ selenophene ratio. Xue et al.528 synthesized a copolymer of indacenodithienothiophene and NDI units (471) to use as an acceptor material in all-polymer solar cells. The photovoltaic properties of this acceptor polymer were investigated using a series donor polymers with a device structure of ITO/ PEDOT:PSS/donor polymer:(471)/PDINO/Al, where power conversion efficiency values up to 5.33% were achieved. Kim et al. described the structure−property relationships associated with a hybrid siloxane-terminated hexyl chain (460, 461), and their influence on molecular packing, thin-film morphology, as well as charge carrier transport were reported for NDI-based polymers bearing NDI and bithiophene repeating units. They found that polymer films cast as a coating from chloroform solvent favor both face-on and edgeon orientation, while 1-chloronaphthalene-cast films only favor an edge-on orientation, which influences the mobility in both of the devices. A mobility of up to 1.04 cm2/V·s was observed for mixed device fabrication from chloroform solvent, which may be due to a highly balanced face-on to edge-on ratio.529 Recently, the Wang group synthesized three new narrowbandgap polymers based on an alkyl-thiophene-flanked NDI (TNDI) unit through direct arylation polycondensation (472− 474). All of the polymers showed high thermal stability up to 335 °C, and among these, the copolymer consisting of alternating TNDI and 4,6-di(2-thienyl)thieno[3,4-c][1,2,5]thiadiazole performed the best as a thin-film transistor, with a moderate hole mobility of 4.6 × 10−3 cm2/V·s under ambient conditions.530 Dai et al. synthesized a series of copolymers, 4,4,9,9-tetrakis(4-hexylphenyl)-indaceno[1,2-b:5,6-b′]-dithiophene (490) as a donor unit, and perylene diimide (PDI) and/ or NDI as an acceptor moiety, by Stille coupling copolymerization. Although the PCEs were not great, it is noteworthy, as Voc is increased from 0.621 to 0.706 V.531 Hou and co-workers first reported a high PCE by incorporating a novel water processed cathode interlayer based on NDI (459b), PCEs of 9.51% were recorded in PBDT-TS1/PC71BM-based polymer photovoltaic cells.532 In another report, they synthesized a novel 2Dconjugated polymer acceptor based on NDI and alkylthiothiophene-substituted benzothophene, and they found that, under the weight ratio of 1.5:1 and 3 vol % of 1-chloronaphthalene, the copolymer-based, all-polymer photovoltaic device exhibited a desirable PCE of nearly 3%.533 Recently, the Kim group described the effect of the alkyl-chain branching position (459a) on the nanoscale morphology of polymers and the performance of all-polymer solar cells.511 Li et al.534 reported

electron transporting layers (ETLs) for use in high performance large-area PSC modules in the future. Analogue polymers to (466 and 467) similar to the system reported Wu et al.518 were introduced by Sun et al.519 as ETL to replace fullerenes in p−i− n planar-heterojunction organometal trihalide perovskite solar cells. They have demonstrated the critical role the amine groups play in improving the interface of ETL/cathode by better work function matching and the interface of perovskite/ETL by more efficient trap passivation, improving all the photovoltaic parameters of the cell. An analogue polymer of (466) without the thiophene moiety (PFTNDI) was investigated by Nam et al.520 in order to improve the surface properties of the ZnO layer and enhance the interfacial contact with the active layer in the inverted ITO/ZnO/PFTNDI/PTB7-Th:PC71BM/MoO3/ Ag configuration polymer solar cells. The dipole moment at the interface of ZnO/PFTNDI reduced the work function of the ITO/ZnO electrode, and significantly decreased the series resistance in the devices. Also they have shown an enhanced uniformity and hydrophobicity of ZnO treated with PFTNDI, improving the contact with the hydrophobic PTB7-Th:PC71BM active layer. Two other reports by Matsidik et al.521 and Senkovskyy et al.522 demonstrated unusual Pd and Ni catalyzed polymerization of NDI anion radicals (459b). These led to very high-performing n-type materials with little structural irregularity and relatively controlled molecular weight. Previous protocols for the synthesis of polymeric n-type semiconductors have often led to limited control of the polymerization process, causing regio-irregularities that have hindered performance. Durban et al.523 synthesized a similar series of compounds containing NDI with a π-spacer consisting of one, two, or three thiopehenes. Their findings revealed an increase in π-spacer between NDI chromophores, increased electron mobility, as well as increased ION/IOFF ratios. All these three studies represent yet again simple design strategies and techniques in order to produce good quality semiconducting materials based on NDI polymers. Structural changes in NDI and other diimide containing polymers are not strictly limited to simple spacers, such as vinyl and thiophene (488 and 468). Zhou et al. reported the synthesis and characterization of an NDI and PDI polymer linked via an N-alkylated phenthiazene. Both materials exhibited broad absorption and displayed typical optical bandgaps of 1.4−1.5 eV. Variances in the molecular weight, which are introduced as N-alkyl substitution of the imide nitrogens, resulted in differences in performance. They showed that the higher the molecular weight, the greater the crystallinity, and uniformity of the film, and the better the π−π-stacking and the better the overall performance as n-type semiconductors. The highest electron mobility value recorded was 0.05 cm2/V·s with an ION/IOFF ratio of 105 under a nitrogen atmosphere.524 Another report by Popere et al. investigated the properties of an interesting dipyrrole, boron, and alkyne containing copolymer (469, 489) with NDI as an acceptor moiety. Synthesis was carried out using a Sonagashira coupling from dibromo-NDI (22) or -PDI, these compounds showed broad absorption with an optical bandgap of 1.63 eV, low laying LUMO levels, and good photosensitivity. OTFT devices made from (469) gave moderate electron mobility values of 3.7 × 10−6 cm2/V·s with an ION/IOFF ratio of 103 and a threshold voltage of approximately 35 V.525 Alvey et al. synthesized an NDI-based polymer via a Stille coupling reaction of dibromoNDI (22) with bis(tributylstannyl)acetylene (Figure 131; 470).526 This synthetic protocol is a very simple way to generate large amounts of material. As expected, strong 11772

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 132. PDI polymers and PDI−NDI copolymers and the role of the spacer on NDI and PDI polymer optoelectronics.

the fabrication of photovoltaic devices with an inverted configuration, with an all polymer active layer composed of (459b) electron acceptor and a series of diketopyrrolopyrrole (DPP) polymers as electron donors. Their results showed the pairing of these polymers gave strong absorption, good energy level alignment, and high crystallinity of DPP polymers, suggesting a system with high performance. The drawback of this system is the reduced crystallinity of the polymer mixture and reduced phase separation, becoming less efficient in generating useful charge carriers. Lee et al.535 studied the effect of N-substitution side chains in (459b) on the performance of all polymer photovoltaic devices. They found the engineering of alkyl side chains of (459b) as n-type polymers can influence their crystalline structures, blend morphology, and performance in OFET devices. More densely ordered thin films were prepared by (459b) with shorter alkyl side chains. On the other hand more thermal annealing has significantly changed the crystallinity of longer side chain-substituted (459b), resulting in the highest electron mobility of 1.90 cm2/ V·s. Anton et al.536

studied the crystallinity of (459b) when spin coated, resulting in various thicknesses using the infrared transition moment orientational analysis technique. They have shown the films consist of platelets of oriented lamellae, composed of inclined molecular subunits self-aggregating in long-range order, which was less affected by the film thickness. The Jenekhe group reported on the design and synthesis of a series of new crystalline semiconducting NDI−selenophene−PDI random copolymers (491) (Figure 132), and they investigated them as electron acceptors with benzodithiophene-thieno[3,4-b]thiophene copolymer as a donor. These BHJ solar cells have a power conversion efficiency (PCE) of 6.3%, Jsc = 18.6 mA/ cm2, and external quantum efficiency = 91%.537 The results clearly demonstrate the crystallinity of a polymer component and thus the compatibility in the blend morphology, and that PCE can be controlled by molecular design in polymer/ polymer blend solar cells. Seferos and co-workers synthesized a series of one donor−two acceptor (D−A1)−(D−A2) random polymers containing a 2,7-carbazole donor and varying 11773

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

compositions of PDI and NDI acceptors (492) via Suzuki coupling polymerization and the study of their optical bandgaps, which ranged from 1.77 to 1.87 eV, depending on acceptor composition.538 Earlier reports based on the NDIthiophene (459b) and NDI-selenophene (475) copolymer acceptors achieved high PCEs of 5.7% and 4.8%, with high electron mobilities of 3.6 × 10−4 and 2.6 × 10−4 cm2/V·s in the blended films, respectively.539,540 The Fu group reported the synthesis of NDI-benzothiadiazole (476) copolymers of PNDIBTH and PNDIBTOC8, varying only in the substitution on the benzothidiazole with two octyloxy groups, and their use as acceptors for the fabrication of all-polymer solar cells (allPSCs). They found that octyloxy substitution resulted in an increase in solubility and molecular weight of the compound, but also altered their optical and electrical properties. The PNDIBTOC8-based acceptor provides an excellent PCE of 3.14% with a high open-circuit voltage (Voc) of 0.90 V, much higher than that of the PNDIBTH-based device, i.e. PCE of 1.20% with Voc of 0.76 V.541 The hole and electron mobility values of the octyloxy group-substituted (476), 4.2 and 2.8 × 10−4 cm2/V·s, respectively, were higher than that of the unsubstituted form, being 1.8 × 10−4 cm2/V·s and 9.4 × 10−5 cm2/V·s, respectively. The Kim and Jenekhe group described the use of poly(NDI-alt-biselenophene) (PNDIBS) (477, 478) as an n-type semiconducting material for development of high performance, nonvolatile, electronic memory devices. The device based on PNDIBS field-effect transistor memory showed excellent charge-trapping and detrapping characteristics, and cycled more than 200 times with a current ratio of 103 between the two binary states with a memory window of 28 V.542 In another report, they have described all-polymer solar cells composed of binary blends of donor poly[4,8-bis(5-(2ethylhexyl)thiophen-2-yl)benzo[1,2-b;4,5-b′]dithiophene-2,6diyl-alt-(4-(2-ethylhexanoyl)-thieno[3,4-b]thiophene-)-2,6diyl] and acceptor polymers naphthalene diimide-selenophene copolymer (PNDIBS) (475) and perylene diimide-selenophene copolymer (PPDIS) (493) that gave PCEs ∼ 2.1% in a blended device, but enhancements of 50%−140% were observed in ternary blend solar cells composed of [PBDTTTCT][PNDIS-HD]1−x[PPDIS]x at 75 wt % PNDIBS in a PCE of 3.2%.543 Goto et al. reported the synthesis of NDI-polymers (459a) using the nonstoichiometric polycondensation approach of Stille coupling polycondensation between 2,5bis(trimethylstannyl)thiophene and dibromo-NDI (21) ranging from 1:1 to 1:10 in the presence of Pd2(dba)3/P(otolyl)3.544 Using this approach, they have synthesized a series of NDIbased semiconducting polymers containing benzodithiophene or dithienopyrrole spacer/donor for ambipolar semiconducting materials, which gave near-infrared absorptions, and these derivatives showed higher electron mobilities of up to 1.5 × 10−2 cm2/V·s.545 They have also reported Negishi-type catalyst transfer polymerization using tBu4ZnLi2 catalysts.546 In an NDI-based polymer where polymerization is conducted via the N-styrenic NDI, a modular assembly of polyNDI as multielectron acceptor and subsequent decoration of the chain terminus by [Ru(2,6-di(quinolin-8-yl)-pyridine)2]2+ dye (494) was synthesized by Schroot et al. (see Figure 133).547 The cyclic voltammetry of the dyad showed two large quasireversible reduction waves at −1.0 and −1.5 V, respectively, and the reversible oxidation of the ruthenium center was detected around +0.5 V. In this dyad, the long-lived excited state of the ruthenium center photosensitizer is exploited to promote efficient charge separation irrespective of the flexible

Figure 133. Chemical structure of NDI-based polymers where monomer units are linked via the N-substitutions on the NDI.

saturated linkage pattern. Prentice et al. reported the synthesis and characterization of another NDI-based polymer that is polymerized using the N-substitution moieties.548 They synthesized a NDI−siloxane series of recyclable polymers (495−498) with high thermal stability. These polymers also showed self-healing properties due to the combination of the simple acid−base-triggered repair system, making them strong candidates for flexible electronic devices. Zhang et al.549 designed and synthesized a multifunctional polymer hybrid consisting of NDI containing polyimide (499) and the πelectron-rich anthracene ended polyhedral oligomer Silsesquioxane (POSS-AN) (500). In this hybrib a charge transfer (CT) complex is formed between the electron-deficient NDI and the electron rich anthracene. The CT interaction between NDI and anthracene can be turned-off through the photodimerization of anthracene using UV light, and then turned-on by heating up to 150 °C. 11774

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 134. Chemical structures of copolymers based on a 1,2,5,6-naphthalenediimide unit.

transport correlation. All of the polymers exhibited excellent electron and hole mobilities, with NDI polymer with four thiophene exhibiting an especially balanced ambipolar charge transport of 0.03 cm2/V·s.553 The Russell group synthesized three conjugated polymer zwitterions of thiophene, diketopyrrolopyrrole (DPP), and naphthalene diimide (NDI) backbones with pendant zwitterions, and they investigated their solar cell efficiency. The thiophene-based copolymer gave PCEs of ∼5%, while the narrow-energy-gap DPP- and NDI-based CPZs performed exceptionally well, giving PCEs of 9.49% and 10.19%, respectively.554 The Luscombe group reported a series of n-type hyperbranched polymers, exhibiting excellent electrochemical stability and variable porosity, and they used these as the cathodes of asymmetric supercapacitors.555 The hyperbranched polymers bear triphenylamine (TPA) cores and NDI terminal units connected through thiophene spacer rings between the TPA and NDI units (482), which are used to alter the porosity. These devices show very good stability over 500 cycles because of stable NDI units, and nitrogen adsorption experiments confirm the thiophene spacers resulted in a concomitant increase in the polymer matrix pore size. Importantly, the porosity of the polymer increases as the diffusion resistance decreases; thus, the charge transfer resistance and equivalent series resistance increase, which is essential for supercapacitor use. The synthesis of semifluoronated N-alkyl-substituted NDI-based polymers (460 and 461) through NDI monomers linked with dithiophene or di(thiophene-2-yl)ethane, with remarkably high electron mobility and rigid backbone organization, was reported by Kang et al.555 Both polymers, with thermal annealing and the introduction of high boiling point solvents, displayed excellent ordering, as determined by GIXD of their thin films, where the electron mobilities for (460 and 461) were shown to be as high as 6.50 and 5.64 cm2/V·s, respectively. Qin et al.556 reported two new NDI polymers (483, 484) linked via a thiophene spacer to a dithiophene-containing donor moiety. Both

The benzothiadiazole−1,2,5,6-NDI donor−acceptor (D−A) polymers (501 and 502) were studied by Zhao et al.,550 showing improved film crystallinity and charge carrier mobility (see Figure 134). They fabricated bottom-gate bottom-contact OFET and photovoltaic cell devices using this set of copolymers and showed improved film crystallinity, morphology, and charge transport characteristics with a record hole mobility of 0.82 cm2/ V·s and On/Off ratio of 105 in comparison to previously reported analogues. Peng’s group developed a type of efficient large band gap (2.0 and 1.98 eV) copolymers of 1,2,5,6-INDI acceptor and benzodithiophene donor (503 and 504) with low lying HOMO below −5.40 eV.551 The XRD analysis of the spin-coated polymers which were utilized to fabricate single-junction polymer solar cells showed that both copolymers had good molecular stacking properties. 9.3. Role of Spacers on NDI Optoelectronics

The role of spacers between donors and NDI acceptors was also evaluated (Figure 131).522−526,552 DeBlase et al.552 designed and synthesized a novel NDI polymer consisting of a core TFP moiety with three connecting NDI subunits (481). This structure leads itself to being “netlike” and, consequently, has a high surface area and porosity. It was demonstrated that with this network,525 when incorporated into porous polymer network (PPN) films, electrochemically generated NDI radicals bind strongly to cations such as K+, Li+, and Mg2+. Thus, such a combination of donor-polymers with the interesting electrical characteristics of NDI-polymer should lead to interesting electrical materials to use in energy storage materials and devices. Szumilo et al. synthesized a series of NDI−thiophene copolymers with one, two, three, or four thiophene units (459a−d) using the Stilles reactions of dibromo-NDI (21) and trimethylstannylthiophene monomers. Furthermore, the effect of the extension of the thiophene donor groups as well as thiophene units of copolymers with two and four thiophene units per monomer was studied in terms of the structure-charge 11775

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Anion−π catalysis plays an important role in enolate chemistry due to its significance in polyketide biosynthesis, which corresponds to carbocation chemistry in steroid and terpenoid biosynthesis, effectively. In this regard Fujisawa et al. studied in detail the ion pair−π interaction that results from either anion−π or cation−π interactions on the same surface of the naphthalene monoimide, as illustrated in Figure 136.563 The main aim of this ion pair−π interaction study was to observe whether both cations and anions interact with one and the same aromatic surface. 1H NMR was used to explore the existence of parallel and antiparallel ion pair−π interactions in covalent systems in the ground state. From this work, parallel ion pair−π interactions were mostly favored, which was later confirmed by computational modeling. This study investigated, and reported in detail, complete and logical results on ion pair−π interactions in covalent and semicovalent systems and gave an excellent platform for future developments in this field. Furthermore, the Matile group investigated carboxylate− guanidinium pairs to explore push−pull chromophores on the surface of 4-amino-1,8-naphthalimides (510), which are covalently positioned, as a better explanation of whether anion−π and cation−π interactions can take place simultaneously on the same aromatic surface (Figure 137).564 These chromophores showed a red-shift of 41 nm in the antiparallel orientation with respect to the push−pull dipole, in less polar solvents, which was close to the theoretical calculations of 70 nm in the ground and excited states with a significant bathochromic effect. Antiparallel ion pair−π interactions result in intramolecular Stark effects that ultimately increase the solvatochromic effect. This was confirmed by aggregation, decreasing shifts with solvent polarity, protonation, and parallel carboxylate−guanidinium pairs. Another feature of carboxylate−guanidinium pairs, obtained from theoretical studies, was interaction with the surfaces of π-acidic naphthalenediimides and π-basic pyrenes.559 Matile’s group further utilized the πacidic surface of a NDI between a proline group and a glutamate group to afford trifunctional catalysts for the stereoselective addition of aldehydes to nitroolefins.565 The nonplanar π surfaces around the chiral sulfoxide connectors have a significant effect on process stereoselectivity. Zhao et al. studied catalysis with anion−π interactions, and they reported experimentally obtained data regarding the contributions of the anion−π interactions, which improved our understanding of organocatalysis.566 This study also enabled us to design novel strategies for better stabilization of anionic transition states. Current data on these modified NDI (511) catalysts containing sulfur indicate that increasing π-acidity results in greater stabilization of the transition state, with considerable Kemp elimination (Figure 138). Initial studies show that anionic tetrahedral intermediates are effectively stabilized by anion−π interactions, for substitution and addition reactions on the carbonyl groups of organic compounds. As compared to cation−π interactions, the newly developed catalysts with anion−π interactions are extremely significant due to their connection with vital principles carried out in nature.276 The Matile group also studied in detail asymmetric anion−π catalysis, which consists of enamination of addition to nitroolefins on π-acidic surfaces. The group successfully performed enamine addition to nitro-olefins, which occurs on the aromatic surface of π-acidic NDIs (511), by placing at one side a proline for enamine formation and at the other side a glutamic acid for nitronate protonation, where the active site is constructed by

polymers showed excellent thermal stability and strong absorption in the visible region and had the desired molecular orbital energy levels for use as a p-type material within a BHJ solar cell. Devices from (483) and (484) were made from TiO2 on ITO glass with PEDOT:PSS. The recorded power conversion efficiencies were 1.94 and 2.36%, respectively. Despite the electron deficiency of NDI, there have been instances in which it was used for p-type material in OTFT devices. Wang et al. reported the use of NDI linked to DPP (485) behaving as an ambipolar material by adjusting the number of π-spacers; in this instance, thiophene spacers were used in between the NDI and DPP units (m = 1, 2). Hole and electron mobilities of PNDI-DPP polymer were well balanced, with values of 5.7 × 10−3 and 1.6 × 10−3 cm2/V·s. This illustrates the possibility for NDI bearing polymers’ use in data storage, image sensors, and programmable logic etc.557

10. CATALYSIS The recently discovered anion−π interaction,558 a noncovalent interaction between an electron-rich anion and an electronpoor aromatic π-system with a strong positive quadrupole moment,559 opened new avenues to the sensing35 and molecular recognition of anions. These anion−π interactions are prominent players in several chemical and biological processes, and they can also be used in designing highly selective ion-receptors and channels.44,560 Among several anions, the interactions of fluoride anion (F−) with π-systems have been comprehensively studied, often using the naphthalenediimide (NDI) fluorophore. NDI is frequently chosen as the π-system, since it can be easily functionalized, can be dissolved in a range of solvents, and has tunable fluorescence properties.58,61,279,561 Kim et al.562 reported a series of watersoluble NDIs (505−509), which recently have found uses in a wide variety of applications, from catalysis, to optoelectronic devices, to biological markers, as shown in Figure 135. Side

Figure 135. Chemical structures of the water-soluble N-alkylsubstituted quaternary ammonium cation at various distances from the NDI core.

chain substitution offers solubility of these materials in a variety of polar and nonpolar solvents. Designing a molecular structure with a positive charge center located at different distances from the NDI core by several methylene units allowed the study of the rate of the base catalyzed hydrolysis reaction of the imide groups in NDI. The ethyl linker based compound (508) hydrolyzed 6.8 times faster than a derivative with a propyl linker compound (509). The equilibrium constants for the reactions with hydroxide showed compound (508) to be less stable than (509) at pH 7.5. Investigating the series (505− 509), they found that the rate of hydrolysis of the NDI increased when the cationic charge is closer to the NDI core. 11776

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 136. Anion−π and cation−π interactions can occur on the same aromatic surface. Reproduced from ref 563. Copyright 2015 American Chemical Society.

Figure 138. Catalysis of the Kemp elimination with anion−π interactions. Reproduced from ref 566. Copyright 2014 American Chemical Society.

Figure 137. Ion pair−π interactions in push−pull chromophores. Reproduced with permission from ref 564. Copyright 2014 WileyVCH.

(511, 512 and 513) (Figure 139).567 The rate of reaction, and its enantioselectivity, gradually increased with increase in the π acidity of the properly trifunctional catalysts. For the first time this group provided evidence for the contribution of anion−π interactions to asymmetric enamine catalysis. Here, the correlation between the π-acidity of stereoisomeric catalysts and diastereoselectivity is much more complex, and is similar to the deep effect of chiral centers right at the edge of their πacidic surface, and also takes into account the significance of rigidified and matched architectures around central π-acidic surfaces. Ongoing research has concentrated on chiral sulfoxides, with bulkier substituents at the edge of the π-acidic surfaces,568 to obtain a better stereoselectivity of the anion−π catalysts, and on determination of their absolute configuration in refined catalysis.569,570

Figure 139. Structure of Wennemers and Jørgensen−Hayashi catalysts (512) and (513), respectively, and possible stabilization of adding enamines (from aldehyde) to nitroolefins (TS1) and nitronate protonation (TS2) on the π-acidic surface of the anion−π-catalyst. Reproduced from ref 567. Copyright 2015 American Chemical Society.

As enolate chemistry is vital in both chemistry and biology, the information obtained is very important and motivates application of anion−π interactions in the field of catalysis to a 11777

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 140. Stepwise oxidation of the core substituents of anion−π tweezers (514) gradually increases the π acidity of the catalyst without global structural changes (515, 516). Reproduced with permission from ref 572. Copyright 2015 the Royal Society of Chemistry.

greater extent.571 For natural product biosynthesis, the addition of anion−π interactions to enolate acceptors is essential. The Matile group also performed studies on the selective acceleration of disfavored enolate addition reactions by anion−π interactions.572 The group reported the role of anion−π interactions in the selectivity of enolate chemistry of malonate half thioesters. For natural product biosynthesis, the addition of anion−π interactions to enolate acceptors is essential. However, these processes generally fail without enzymes, due to nonproductive decarboxylation, which takes place predominately. They have also designed and synthesized novel anion−π tweezers (514−516), which accelerate the disfavored and decelerate the favored process by reversing the selectivity (Figures 140 and 141).572 This work clearly indicates

reported important facts regarding enolates, which can be stabilized by up to two pKa units by single unoptimized anion−π interactions. When enones and nitro-olefins are aided by these anion−π stabilized reactive enolate intermediates, this results in the stabilization of the transition-state at approximately 11 kJ/mol, and simultaneously increases the rate of anionic cascade reactions on π-acidic surfaces (Figure 142). In 2015, the Parquette group reported on studies of selfassembled NDI nanotubes.161 The NDI, bearing proline−lysine dipeptide (529), was self-assembled in the presence of a substrate under aqueous conditions (Figure 143) and was subsequently used to catalyze simple aldol condensation reactions in water. The nanotubes provided a hydrophobic microenvironment in which the reactions could occur. Interestingly, enantiomeric excess increased, after five cycles, from 83.5 to 93.5%. This simple system illustrates the use of self-assembled structures having “enzyme-like” properties and is a key step forward in the production of efficient catalysts for chemical reactions. The Matile group’s recent work studies the anion−π catalysis of enolate chemistry, where rigidified Leonard turns are specifically designed as a general motif to precisely position the substrate, to facilitate reactions on aromatic surfaces. They concluded that precisely engineered turns enable efficient anion−π catalysis to take place, even with relatively weak πacidity. Using an electron-deficient aromatic surface, anion−π catalysis occurs efficiently even when substrate placement is less than ideal but selectivity toward the unfavored product increases with increased rigidity of the Leonard turns and directs the substrate molecule in a less directional anion−π interaction environment.574

Figure 141. Structure of substrates (517−520), favored product (521), disfavored product (522), the minimalist bifunctional catalyst (523), anion−π tweezer (524), and control bases (525) and (526). Reproduced with permission from ref 572. Copyright 2015 the Royal Society of Chemistry.

11. CONCLUSION AND OUTLOOK Ample structural design opportunities, well-defined morphologies, simplistic device fabrication, and excellent optoelectronic performances make naphthalene diimides (NDIs) very promising in a diverse array of fields. Imide and coresubstituted NDIs have further added to the perceived use of these materials, owing to their colorful appearance and charge mobilities, within precise supramolecular organizations. This review details the achievements over the past decade in the field of naphthalene diimides, and specifically their application to supramolecular chemistry, sensors, and host− guest complexes for molecular switching devices such as catenanes and rotaxanes, ion-channels by ligand gating, gelators for sensing aromatic systems, catalysis through anion−π

that anion−π interactions have a crucial, contributory role by increasing the selectivity and increasing the π acidity of the catalyst, simultaneously. Recently, Zhao et al. reported a brief account of enolate chemistry, with anion−π interactions which occur on the surface of π-acidic NDI (528) planes with positive quadrupole moments.573 They also verified their crucial role in the binding and transport of anions, and their contribution to the stabilization of anionic reactive intermediates. They have also 11778

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Figure 143. Self-assembly of Pro−Lys dipeptide (529) into nanotubes, and the nanotube-catalyzed aldol reaction. Reproduced with permission from ref 161. Copyright 2015 the Royal Society of Chemistry.

abilities are still uncommon. Research on well-designed NDI architectures will in any case contribute to the major effort in basic research on smart optoelectric nanomaterials that are needed to meet tomorrow’s energy demands in a sustainable way. Future goals and prospects will still be centered on these aspects. Moreover, in structural design, the cooperation of “bottom-up” self-assembly (1D and 3D), together with conventional “top-down” techniques and novel nanoscale manufacturing, may offer new ways to meet the challenge of real device development. Research into the use of 1D and 3D NDI assembled nanotubular structures in practical devices is currently in progress.161 Relevant techniques for fabrication and morphologically controlled assembly should also provide important information for other molecular architectures that bridge the gap between the molecular and macroscopic worlds. From a fundamental viewpoint, molecularly well-defined π-conjugated oligomers remain ideal model systems for the investigation of the relationships between photon−electron conversion and long-lived charge separation.

Figure 142. Positioning and stabilization of enolate anions on π-acidic surfaces as demonstrated by 1H NMR spectra of a 1:1 mixture of (527) and (528) in CDCl3 in the presence of increasing amounts of 1,1,3,3-tetramethylguanidine show that (528) is more acidic than (527), that is, a stabilization of the enolate anion on the π-acidic surface of (528). Reproduced with permission from ref 573. Copyright 2014 Nature PNG publisher.

AUTHOR INFORMATION Corresponding Author

interactions, and NDI intercalations with DNA for medicinal applications. Due to the renowned optical activity of NDIs, and the ease with which their redox properties can be modified, particular, and warranted, attention has been paid within to the use of NDIs in artificial photosynthetic and optoelectronic devices. However, although significant advances have been made, current materials still fall short of the required level for commercial application in practical optoelectronics, and an indepth appreciation of the microcosmic and dynamic selfassembly process is required in order to bring about desired further increases in device performance. In this regard, a large amount of research has focused on morphological control and property optimization through core-substitution of NDIs, but very little work has been carried out on the use of coresubstituted NDIs in real world applications, such as cell-labeling etc. In view of new solution-dispersible applications, for example photocatalytic systems, optimized NDIs/semiconductor pairs with effective binding force and interfacial charge separation

*E-mail: [email protected]; bsheshanath@gmail. com. Notes

The authors declare no competing financial interest. Biographies Mohammad Al Kobaisi, born in 1971 in Iraq, was awarded his M.Sc. in Physical Chemistry from Shahid Beheshti University in 1999, and his Ph.D. under the supervision of A/Prof. Colin Rix and Prof. David Mainwaring from RMIT University in 2007, investigating the design of new selective adsorbent polymers for sensing applications. Since 2007 he has continued his research career working in Prof. David Mainwaring’s team in designing drug delivery and vaccine delivery systems in collaboration with the pharmaceutical industry. His general research interest remains in molecular design and in developing nanoand microfunctional materials as vaccine delivery systems. Sidhanath V. Bhosale, is a native of Indral (Udgir), India. Dr. Bhosale worked as a Lecturer at S.RT.M.U. Nanded and K.E.S. Anandibai 11779

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

REFERENCES

Pradhan Science College, Nagothane, Raigad, before moving to FU Berlin, Germany, where he received his doctorate degree in chemistry under the supervision of Prof. J. H. Fuhrhop. He undertook postdoctoral study under the supervision of Dr. Kelly Velonia at the University of Geneva, Switzerland. He has worked as Reader and then as Associate Professor at North Maharashtra University, Jalgaon, for five years. Currently, he is working at the Polymers and Functional Materials Division, CSIR-IICT, Hyderabad, as a Senior Scientist. His research interests are in the areas of supramolecular chemistry and molecular materials.

(1) Bhosale, S. V.; Jani, C. H.; Langford, S. J. Chemistry of Naphthalene Diimides. Chem. Soc. Rev. 2008, 37 (2), 331−342. (2) Würthner, F. Perylene Bisimide Dyes as Versatile Building Blocks for Functional Supramolecular Architectures. Chem. Commun. 2004, 1564−1579. (3) Huang, C.; Barlow, S.; Marder, S. R. Perylene-3,4,9,10tetracarboxylic Acid Diimides: Synthesis, Physical Properties, and Use in Organic Electronics. J. Org. Chem. 2011, 76 (8), 2386−2407. (4) Würthner, F.; Stolte, M. Naphthalene and Perylene Diimides for Organic Transistors. Chem. Commun. 2011, 47 (18), 5109−5115. (5) Zhan, X.; Facchetti, A.; Barlow, S.; Marks, T. J.; Ratner, M. A.; Wasielewski, M. R.; Marder, S. R. Rylene and Related Diimides for Organic Electronics. Adv. Mater. 2011, 23 (2), 268−284. (6) Jones, B. A.; Facchetti, A.; Wasielewski, M. R.; Marks, T. J. Tuning Orbital Energetics in Arylene Diimide Semiconductors. Materials Design for Ambient Stability of n-Type Charge Transport. J. Am. Chem. Soc. 2007, 129 (49), 15259−15278. (7) Herbst, W.; Hunger, K.; Wilker, G.; Ohleier, H.; Winter, R. In Industrial Organic Pigments; Wiley-VCH Verlag GmbH & Co. KGaA: 2005. (8) Langhals, H. Cyclic Carboxylic Imide Structures as Structure Elements of High Stability. Novel Developments in Perylene Dye Chemistry. Heterocycles 1995, 40 (1), 477−500. (9) Holtrup, F. O.; Müller, G. R. J.; Quante, H.; De Feyter, S.; De Schryver, F. C.; Müllen, K. Terrylenimides: New NIR Fluorescent Dyes. Chem. - Eur. J. 1997, 3 (2), 219−225. (10) Quante, H.; Müllen, K. Quaterrylenebis(dicarboximides). Angew. Chem., Int. Ed. Engl. 1995, 34 (12), 1323−1325. (11) Avlasevich, Y.; Li, C.; Müllen, K. Synthesis and Applications of Core-Enlarged Perylene Dyes. J. Mater. Chem. 2010, 20 (19), 3814− 3826. (12) De Schryver, F. C.; Vosch, T.; Cotlet, M.; Van der Auweraer, M.; Müllen, K.; Hofkens, J. Energy Dissipation in Multichromophoric Single Dendrimers. Acc. Chem. Res. 2005, 38 (7), 514−522. (13) Weil, T.; Vosch, T.; Hofkens, J.; Peneva, K.; Müllen, K. The Rylene Colorant Family-Tailored Nanoemitters for Photonics Research and Applications. Angew. Chem., Int. Ed. 2010, 49 (48), 9068−9093. (14) Liu, Z.; Zhang, G.; Cai, Z.; Chen, X.; Luo, H.; Li, Y.; Wang, J.; Zhang, D. New Organic Semiconductors with Imide/AmideContaining Molecular Systems. Adv. Mater. 2014, 26 (40), 6965− 6977. (15) Castellano, F. N. Transition Metal Complexes Meet the Rylenes. Dalton Trans. 2012, 41 (28), 8493−8501. (16) Würthner, F. Bay-Substituted Perylene Bisimides: Twisted Fluorophores for Supramolecular Chemistry. Pure Appl. Chem. 2006, 78 (12), 2341−2349. (17) Görl, D.; Zhang, X.; Würthner, F. Molecular Assemblies of Perylene Bisimide Dyes in Water. Angew. Chem., Int. Ed. 2012, 51 (26), 6328−6348. (18) Würthner, F.; Saha-Möller, C. R.; Fimmel, B.; Ogi, S.; Leowanawat, P.; Schmidt, D. Perylene Bisimide Dye Assemblies as Archetype Functional Supramolecular Materials. Chem. Rev. 2016, 116 (3), 962−1052. (19) Guha, S.; Goodson, F. S.; Corson, L. J.; Saha, S. Boundaries of Anion/Naphthalenediimide Interactions: From Anion−π Interactions to Anion-Induced Charge-Transfer and Electron-Transfer Phenomena. J. Am. Chem. Soc. 2012, 134 (33), 13679−13691. (20) Katz, H. E.; Lovinger, A. J.; Johnson, J.; Kloc, C.; Siegrist, T.; Li, W.; Lin, Y. Y.; Dodabalapur, A. A Soluble and Air-stable Organic Semiconductor with High Electron Mobility. Nature 2000, 404 (6777), 478−481. (21) Suraru, S.-L.; Würthner, F. Strategies for the Synthesis of Functional Naphthalene Diimides. Angew. Chem., Int. Ed. 2014, 53 (29), 7428−7448. (22) Thalacker, C.; Röger, C.; Würthner, F. Synthesis and Optical and Redox Properties of Core-Substituted Naphthalene Diimide Dyes. J. Org. Chem. 2006, 71 (21), 8098−8105.

Kay Latham, born in 1964, completed her Ph.D. (isomorphous substitution of zeolites) in the United Kingdom, under the supervision of Craig D. Williams. Prior to this, she had a successful career in the UK surface coatings industry. Following her Ph.D., she undertook a higher degree in Education at Cambridge University (Sidney Sussex College), and postdoctoral studies sponsored by British Gas. She moved to Australia as a junior academic in 1999. Now Head of Chemistry and Environmental Science at RMIT University, Melbourne, and President of the RACI in Victoria, she has published widely in the area of materials and sensor development. Aaron Raynor, born in 1989, is a current Ph.D. student of Dr. Sheshanath Bhosale at RMIT University. He completed his B.Sc. (Hons) in 2011, having worked on the synthesis of novel nitro-olefins for use as antimicrobial agents. During his Ph.D., he has worked and published in the area of organic photovoltaics, chemical sensing, and self-assembly. His research interests are predominately in the area of organic synthesis and electronics, with his Ph.D. being on the synthesis and application of small molecules for use in bulk heterojunction organic solar cells. Sheshanath V Bhosale (Shesh), born in 1976 in Indral Tq. Devoni, completed his M.Sc. in chemistry from the Udgiri college Udgir (S.R.T.M.U. Nanded) in 1999. He then worked as a project assistant at NCL, Pune, before moving to the Freie University Berlin, Germany, where he received his Ph.D. (Magna Cum Lauda) in supramolecular chemistry under the supervision of Prof. J. H. Fuhrhop in 2004. Dr. Bhosale pursued his postdoctoral studies with Prof. S. Matile at University of Geneva, Switzerland, under the auspices of a Roche Foundation Fellowship. This was followed by a stay at Monash University, Australia, for five years as an ARC-APD Fellow. Shesh received a prestigious Future Fellowship (2011) from the Australian Research Council and is now working at the School of Science, RMIT University, Melbourne, Australia. Shesh is interested in design and synthesis of π-functional materials, especially small molecules, for sensing, biomaterials and supramolecular chemistry applications.

ACKNOWLEDGMENTS S.V.B. (RMIT) acknowledges financial support from the Australian Research Council (ARC), Australia, under a Future Fellowship Scheme (FT110100152). S.V.B. (IICT) is grateful for financial support from the Intelcoat project CSC0114, CSIR, DAE-BRNS (Project Code: 37(2)/14/08/2014-BRNS), Mumbai, and SERB (DST) SB/S1/IC-09/2014 India. The School of Science, RMIT University, is acknowledged for providing support through a Visiting Fellow position for M.AlK. We also acknowledge all the current and past group members for their contributions in the NDI field and B. Alford and P. Sonar for their help in the outlined synthesis, medicinal, and solar cell parts, respectively. 11780

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

(23) Chen, Z.; Lohr, A.; Saha-Möller, C. R.; Würthner, F. SelfAssembled π-Stacks of Functional Dyes in Solution: Structural and Thermodynamic Features. Chem. Soc. Rev. 2009, 38 (2), 564−584. (24) Miller, L. L.; Mann, K. R. π-Dimers and π-Stacks in Solution and in Conducting Polymers. Acc. Chem. Res. 1996, 29 (9), 417−423. (25) Debreczeny, M. P.; Svec, W. A.; Wasielewski, M. R. Optical Control of Photogenerated Ion Pair Lifetimes: An Approach to a Molecular Switch. Science 1996, 274 (5287), 584−587. (26) Borgström, M.; Shaikh, N.; Johansson, O.; Anderlund, M. F.; Styring, S.; Åkermark, B.; Magnuson, A.; Hammarström, L. Light Induced Manganese Oxidation and Long-Lived Charge Separation in a Mn2II,II−RuII(bpy)3−Acceptor Triad. J. Am. Chem. Soc. 2005, 127 (49), 17504−17515. (27) Chaignon, F.; Buchet, F.; Blart, E.; Falkenstrom, M.; Hammarstrom, L.; Odobel, F. Photoinduced Electron Transfer in Ruthenium(II) Trisbipyridine Complexes Connected to a Naphthalenebisimide via an Oligo(phenyleneethynylene) Spacer. New J. Chem. 2009, 33 (2), 408−416. (28) Iengo, E.; Pantos, G. D.; Sanders, J. K. M.; Orlandi, M.; Chiorboli, C.; Fracasso, S.; Scandola, F. A Fully Self-Assembled NonSymmetric Triad for Photoinduced Charge Separation. Chem. Sci. 2011, 2 (4), 676−685. (29) Gunderson, V. L.; Smeigh, A. L.; Kim, C. H.; Co, D. T.; Wasielewski, M. R. Electron Transfer within Self-Assembling Cyclic Tetramers Using Chlorophyll-Based Donor−Acceptor Building Blocks. J. Am. Chem. Soc. 2012, 134 (9), 4363−4372. (30) Natali, M.; Ravaglia, M.; Scandola, F.; Boixel, J.; Pellegrin, Y.; Blart, E.; Odobel, F. Long-Range Charge Separation in a Ferrocene− (Zinc Porphyrin)−Naphthalenediimide Triad. Asymmetric Role of 1,2,3-Triazole Linkers. J. Phys. Chem. C 2013, 117 (38), 19334−19345. (31) Volker, S. F.; Schmiedel, A.; Holzapfel, M.; Bohm, C.; Lambert, C. Charge Transfer Dynamics in Squaraine-Naphthalene Diimide Copolymers. Phys. Chem. Chem. Phys. 2013, 15 (45), 19831−19844. (32) Vignon, S. A.; Jarrosson, T.; Iijima, T.; Tseng, H.-R.; Sanders, J. K. M.; Stoddart, J. F. Switchable Neutral Bistable Rotaxanes. J. Am. Chem. Soc. 2004, 126 (32), 9884−9885. (33) Dey, S. K.; Coskun, A.; Fahrenbach, A. C.; Barin, G.; Basuray, A. N.; Trabolsi, A.; Botros, Y. Y.; Stoddart, J. F. A Redox-Active Reverse Donor-Acceptor Bistable [2]Rotaxane. Chem. Sci. 2011, 2 (6), 1046− 1053. (34) Scott Lokey, R.; Iverson, B. L. Synthetic Molecules that Fold into a Pleated Secondary Structure in Solution. Nature 1995, 375 (6529), 303−305. (35) Talukdar, P.; Bollot, G.; Mareda, J.; Sakai, N.; Matile, S. Synthetic Ion Channels with Rigid-Rod π-Stack Architecture that Open in Response to Charge-Transfer Complex Formation. J. Am. Chem. Soc. 2005, 127 (18), 6528−6529. (36) Mitra, A.; Hubley, C. T.; Panda, D. K.; Clark, R. J.; Saha, S. Anion-Directed Assembly of a Non-Interpenetrated Square-Grid Metal-Organic Framework with Nanoscale Porosity. Chem. Commun. 2013, 49 (59), 6629−6631. (37) Jin, S.; Ding, X.; Feng, X.; Supur, M.; Furukawa, K.; Takahashi, S.; Addicoat, M.; El-Khouly, M. E.; Nakamura, T.; Irle, S.; Fukuzumi, S.; Nagai, A.; Jiang, D. Charge Dynamics in A Donor−Acceptor Covalent Organic Framework with Periodically Ordered Bicontinuous Heterojunctions. Angew. Chem., Int. Ed. 2013, 52 (7), 2017−2021. (38) Avinash, M. B.; Govindaraju, T. Amino Acid Derivatized Arylenediimides: A Versatile Modular Approach for Functional Molecular Materials. Adv. Mater. 2012, 24 (29), 3905−3922. (39) Ponnuswamy, N.; Pantoş, G. D.; Smulders, M. M. J.; Sanders, J. K. M. Thermodynamics of Supramolecular Naphthalenediimide Nanotube Formation: The Influence of Solvents, Side Chains, and Guest Templates. J. Am. Chem. Soc. 2012, 134 (1), 566−573. (40) Pantoş, G. D.; Wietor, J.-L.; Sanders, J. K. M. Filling Helical Nanotubes with C60. Angew. Chem., Int. Ed. 2007, 46 (13), 2238− 2240. (41) Mukhopadhyay, P.; Iwashita, Y.; Shirakawa, M.; Kawano, S.-i.; Fujita, N.; Shinkai, S. Spontaneous Colorimetric Sensing of the

Positional Isomers of Dihydroxynaphthalene in a 1D Organogel Matrix. Angew. Chem., Int. Ed. 2006, 45 (10), 1592−1595. (42) Stewart, W. W. Lucifer Dyes-Highly Fluorescent Dyes for Biological Tracing. Nature 1981, 292 (5818), 17−21. (43) Mareda, J.; Matile, S. Anion−π Slides for Transmembrane Transport. Chem. - Eur. J. 2009, 15 (1), 28−37. (44) Dawson, R. E.; Hennig, A.; Weimann, D. P.; Emery, D.; Ravikumar, V.; Montenegro, J.; Takeuchi, T.; Gabutti, S.; Mayor, M.; Mareda, J.; Schalley, C. A.; Matile, S. Experimental Evidence for the Functional Relevance of Anion−π Interactions. Nat. Chem. 2010, 2 (7), 533−538. (45) Míšek, J.; Vargas Jentzsch, A.; Sakurai, S.-i.; Emery, D.; Mareda, J.; Matile, S. A Chiral and Colorful Redox Switch: Enhanced π Acidity in Action. Angew. Chem., Int. Ed. 2010, 49 (42), 7680−7683. (46) Zhao, Y.; Domoto, Y.; Orentas, E.; Beuchat, C.; Emery, D.; Mareda, J.; Sakai, N.; Matile, S. Catalysis with Anion−π Interactions. Angew. Chem., Int. Ed. 2013, 52 (38), 9940−9943. (47) Guha, S.; Saha, S. Fluoride Ion Sensing by an Anion−π Interaction. J. Am. Chem. Soc. 2010, 132 (50), 17674−17677. (48) Guha, S.; Goodson, F. S.; Roy, S.; Corson, L. J.; Gravenmier, C. A.; Saha, S. Electronically Regulated Thermally and Light-Gated Electron Transfer from Anions to Naphthalenediimides. J. Am. Chem. Soc. 2011, 133 (39), 15256−15259. (49) Gabriel, G. J.; Iverson, B. L. Aromatic Oligomers that Form Hetero Duplexes in Aqueous Solution. J. Am. Chem. Soc. 2002, 124 (51), 15174−15175. (50) Hamilton, D. G.; Davies, J. E.; Prodi, L.; Sanders, J. K. M. Synthesis, Structure and Photophysics of Neutral π-Associated [2]Catenanes. Chem. - Eur. J. 1998, 4 (4), 608−620. (51) Hamilton, D. G.; Montalti, M.; Prodi, L.; Fontani, M.; Zanello, P.; Sanders, J. K. M. Photophysical and Electrochemical Characterisation of the Interactions between Components in Neutral πAssociated [2]Catenanes. Chem. - Eur. J. 2000, 6 (4), 608−617. (52) Cougnon, F. B. L.; Jenkins, N. A.; Pantoş, G. D.; Sanders, J. K. M. Templated Dynamic Synthesis of a [3]Catenane. Angew. Chem., Int. Ed. 2012, 51 (6), 1443−1447. (53) Iijima, T.; Vignon, S. A.; Tseng, H.-R.; Jarrosson, T.; Sanders, J. K. M.; Marchioni, F.; Venturi, M.; Apostoli, E.; Balzani, V.; Stoddart, J. F. Controllable Donor−Acceptor Neutral [2]Rotaxanes. Chem. - Eur. J. 2004, 10 (24), 6375−6392. (54) Brouwer, A. M.; Frochot, C.; Gatti, F. G.; Leigh, D. A.; Mottier, L. c.; Paolucci, F.; Roffia, S.; Wurpel, G. W. H. Photoinduction of Fast, Reversible Translational Motion in a Hydrogen-Bonded Molecular Shuttle. Science 2001, 291 (5511), 2124−2128. (55) Ponnuswamy, N.; Cougnon, F. B. L.; Clough, J. M.; Pantoş, G. D.; Sanders, J. K. M. Discovery of an Organic Trefoil Knot. Science 2012, 338 (6108), 783−785. (56) Molla, M. R.; Das, A.; Ghosh, S. Self-Sorted Assembly in a Mixture of Donor and Acceptor Chromophores. Chem. - Eur. J. 2010, 16 (33), 10084−10093. (57) Das, A.; Molla, M. R.; Banerjee, A.; Paul, A.; Ghosh, S. Hydrogen-Bonding Directed Assembly and Gelation of Donor− Acceptor Chromophores: Supramolecular Reorganization from a Charge-Transfer State to a Self-Sorted State. Chem. - Eur. J. 2011, 17 (22), 6061−6066. (58) Bhosale, S. V.; Bhosale, S. V.; Bhargava, S. K. Recent Progress of Core-Substituted Naphthalenediimides: Highlights from 2010. Org. Biomol. Chem. 2012, 10 (32), 6455−6468. (59) Vollmann, H.; Becker, H.; Corell, M.; Streeck, H. Beiträge zur Kenntnis des Pyrens und seiner Derivate. Justus Liebigs Ann. Chem. 1937, 531 (1), 1−159. (60) Würthner, F.; Ahmed, S.; Thalacker, C.; Debaerdemaeker, T. Core-Substituted Naphthalene Bisimides: New Fluorophors with Tunable Emission Wavelength for FRET Studies. Chem. - Eur. J. 2002, 8 (20), 4742−4750. (61) Sakai, N.; Mareda, J.; Vauthey, E.; Matile, S. Core-Substituted Naphthalenediimides. Chem. Commun. 2010, 46 (24), 4225−4237. (62) Bhosale, S.; Sisson, A. L.; Talukdar, P.; Fürstenberg, A.; Banerji, N.; Vauthey, E.; Bollot, G.; Mareda, J.; Röger, C.; Würthner, F.; Sakai, 11781

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

N.; Matile, S. Photoproduction of Proton Gradients with π-Stacked Fluorophore Scaffolds in Lipid Bilayers. Science 2006, 313 (5783), 84− 86. (63) Chopin, S.; Chaignon, F.; Blart, E.; Odobel, F. Syntheses and Properties of Core-Substituted Naphthalene Bisimides with Aryl Ethynyl or Cyano Groups. J. Mater. Chem. 2007, 17 (39), 4139−4146. (64) Krüger, H.; Janietz, S.; Sainova, D.; Dobreva, D.; Koch, N.; Vollmer, A. Hybrid Supramolecular Naphthalene Diimide-thiophene Structures and their Application in Polymer Electronics. Adv. Funct. Mater. 2007, 17 (18), 3715−3723. (65) Suraru, S.-L.; Würthner, F. Core-Tetrasubstituted Naphthalene Diimides by Stille Cross-Coupling Reactions and Characterization of Their Optical and Redox Properties. Synthesis 2009, 2009 (11), 1841− 1845. (66) Guo, X.; Watson, M. D. Conjugated Polymers from Naphthalene Bisimide. Org. Lett. 2008, 10 (23), 5333−5336. (67) Jones, B. A.; Facchetti, A.; Marks, T. J.; Wasielewski, M. R. Cyanonaphthalene Diimide Semiconductors for Air-Stable, Flexible, and Optically Transparent n-Channel Field-Effect Transistors. Chem. Mater. 2007, 19 (11), 2703−2705. (68) Röger, C.; Würthner, F. Core-Tetrasubstituted Naphthalene Diimides: Synthesis, Optical Properties, and Redox Characteristics. J. Org. Chem. 2007, 72 (21), 8070−8075. (69) Röger, C.; Ahmed, S.; Würthner, F. Synthesis and Optical Properties of Fluorescent Core-Trisubstituted Naphthalene Diimide Dyes. Synthesis 2007, 2007 (12), 1872−1876. (70) Gao, X.; Qiu, W.; Yang, X.; Liu, Y.; Wang, Y.; Zhang, H.; Qi, T.; Liu, Y.; Lu, K.; Du, C.; Shuai, Z.; Yu, G.; Zhu, D. First Synthesis of 2,3,6,7-Tetrabromonaphthalene Diimide. Org. Lett. 2007, 9 (20), 3917−3920. (71) Sakai, N.; Sisson, A. L.; Bürgi, T.; Matile, S. Zipper Assembly of Photoactive Rigid-Rod Naphthalenediimide π-Stack Architectures on Gold Nanoparticles and Gold Electrodes. J. Am. Chem. Soc. 2007, 129 (51), 15758−15759. (72) Kishore, R. S. K.; Kel, O.; Banerji, N.; Emery, D.; Bollot, G.; Mareda, J.; Gomez-Casado, A.; Jonkheijm, P.; Huskens, J.; Maroni, P.; Borkovec, M.; Vauthey, E.; Sakai, N.; Matile, S. Ordered and Oriented Supramolecular n/p-Heterojunction Surface Architectures: Completion of the Primary Color Collection. J. Am. Chem. Soc. 2009, 131 (31), 11106−11116. (73) Sakai, N.; Bhosale, R.; Emery, D.; Mareda, J.; Matile, S. Supramolecular n/p-Heterojunction Photosystems with Antiparallel Redox Gradients in Electron- and Hole-Transporting Pathways. J. Am. Chem. Soc. 2010, 132 (20), 6923−6925. (74) Pugliesi, I.; Krok, P.; Lochbrunner, S.; Błaszczyk, A.; von Hänisch, C.; Mayor, M.; Riedle, E. Variation of the Ultrafast Fluorescence Quenching in 2,6-Sulfanyl-Core-Substituted Naphthalenediimides by Electron Transfer. J. Phys. Chem. A 2010, 114 (48), 12555−12560. (75) Pugliesi, I.; Megerle, U.; Suraru, S.-L.; Würthner, F.; Riedle, E.; Lochbrunner, S. Influence of Core-Substitution on the Ultrafast Charge Separation and Recombination in Arylamino Core-Substituted Naphthalene Niimides. Chem. Phys. Lett. 2011, 504 (1−3), 24−28. (76) Grunder, S.; Muñoz Torres, D.; Marquardt, C.; Błaszczyk, A.; Krupke, R.; Mayor, M. Synthesis and Optical Properties of Molecular Rods Comprising a Central Core-Substituted Naphthalenediimide Chromophore for Carbon Nanotube Junctions. Eur. J. Org. Chem. 2011, 2011 (3), 478−496. (77) Marquardt, C. W.; Grunder, S.; Blaszczyk, A.; Dehm, S.; Hennrich, F.; Lohneysen, H. v.; Mayor, M.; Krupke, R. Electroluminescence from a Single Nanotube-Molecule-Nanotube Junction. Nat. Nanotechnol. 2010, 5 (12), 863−867. (78) Horne, W. S.; Ashkenasy, N.; Ghadiri, M. R. Modulating Charge Transfer through Cyclic D,L-α-Peptide Self-Assembly. Chem. - Eur. J. 2005, 11 (4), 1137−1144. (79) Pengo, P.; Pantoş, G. D.; Otto, S.; Sanders, J. K. M. Efficient and Mild Microwave-Assisted Stepwise Functionalization of Naphthalenediimide with α-Amino Acids. J. Org. Chem. 2006, 71 (18), 7063−7066.

(80) Tambara, K.; Ponnuswamy, N.; Hennrich, G.; Pantoş, G. D. Microwave-Assisted Synthesis of Naphthalenemonoimides and NDesymmetrized Naphthalenediimides. J. Org. Chem. 2011, 76 (9), 3338−3347. (81) Kolhe, N. B.; Ashar, A. Z.; Narayan, K. S.; Asha, S. K. Naphthalene Diimide Copolymers with Oligo(p-phenylenevinylene) and Benzobisoxazole for Balanced Ambipolar Charge Transport. Macromolecules 2014, 47 (7), 2296−2305. (82) Rauch, G.; Hoger, S. N-Alkylated and N,N-dialkylated 1,6diaminoperylene diimides synthesized via copper catalyzed direct aromatic amination. Chem. Commun. 2014, 50 (42), 5659−5661. (83) Wu, Y.; Frasconi, M.; Gardner, D. M.; McGonigal, P. R.; Schneebeli, S. T.; Wasielewski, M. R.; Stoddart, J. F. Electron Delocalization in a Rigid Cofacial Naphthalene-1,8:4,5-bis(dicarboximide) Dimer. Angew. Chem. 2014, 126 (36), 9630−9635. (84) Sakai, N.; Matile, S. Multistep Organic Synthesis of Modular Photosystems. Beilstein J. Org. Chem. 2012, 8, 897−904. (85) Leng, B.; Lu, D.; Jia, X.; Yang, X.; Gao, X. Synthesis of Monolateral and Bilateral Sulfur-Heterocycle Fused Naphthalene Diimides (NDIs) from Monobromo and Dibromo NDIs. Org. Chem. Front. 2015, 2 (4), 372−377. (86) Fukutomi, Y.; Nakano, M.; Hu, J.-Y.; Osaka, I.; Takimiya, K. Naphthodithiophenediimide (NDTI): Synthesis, Structure, and Applications. J. Am. Chem. Soc. 2013, 135 (31), 11445−11448. (87) Li, C.; Xiao, C.; Li, Y.; Wang, Z. Synthesis and Properties of Heterocyclic Acene Diimides. Org. Lett. 2013, 15 (3), 682−685. (88) Cai, K.; Xie, J.; Zhao, D. NIR J-Aggregates of Hydroazaheptacene Tetraimides. J. Am. Chem. Soc. 2014, 136 (1), 28−31. (89) Li, C.; Jiang, W.; Zhu, X.; Wang, Z. Synthesis and Properties of Diazapentacene Diimides. Asian J. Org. Chem. 2014, 3 (2), 114−117. (90) Suraru, S.-L.; Burschka, C.; Würthner, F. Diindole-Annulated Naphthalene Diimides: Synthesis and Optical and Electronic Properties of Syn- and Anti-Isomers. J. Org. Chem. 2014, 79 (1), 128−139. (91) Chen, X.; Zhang, G.; Luo, H.; Li, Y.; Liu, Z.; Zhang, D. Ambipolar Charge-Transport Property for the D-A Complex with Naphthalene Diimide Motif. J. Mater. Chem. C 2014, 2 (16), 2869− 2876. (92) Cai, K.; Xie, J.; Yang, X.; Zhao, D. Heterohexacene Diimides: Anti- and Syn- Isomers and Quinonoid Forms. Org. Lett. 2014, 16 (7), 1852−1855. (93) Berezin, A. A.; Sciutto, A.; Demitri, N.; Bonifazi, D. Rational Synthesis of AB-Type N-Substituted Core-Functionalized Naphthalene Diimides (cNDIs). Org. Lett. 2015, 17 (8), 1870−1873. (94) Maniam, S.; Sandanayake, S.; Izgorodina, E. I.; Langford, S. J. Unusual Products from Oxidation of Naphthalene Diimides. Asian J. Org. Chem. 2016, 5, 490. (95) Bijak, K.; Filapek, M.; Wiacek, M.; Janeczek, H.; Grucela, M.; Smolarek, K.; Mackowski, S.; Schab-Balcerzak, E. Preparation and Characterization of new Aliphatic-Tailed five- and six-membered Azomethine-Diimides. Mater. Chem. Phys. 2016, 171, 97−108. (96) Matsidik, R.; Komber, H.; Sommer, M. Rational Use of Aromatic Solvents for Direct Arylation Polycondensation: C−H Reactivity versus Solvent Quality. ACS Macro Lett. 2015, 4 (12), 1346−1350. (97) Ke, H.; Jiao, C.; Qian, Y.-H.; Lin, M.-J.; Chen, J.-Z. Naphthalene Diimide Templated Synthesis of Pillar[6]arenes. Chin. J. Chem. 2015, 33 (3), 339−342. (98) Lehn, J.-M. In Supramol. Chem.; Wiley-VCH Verlag GmbH & Co. KGaA: 2006. (99) Whitesides, G. M.; Simanek, E. E.; Mathias, J. P.; Seto, C. T.; Chin, D.; Mammen, M.; Gordon, D. M. Noncovalent Synthesis: Using Physical-Organic Chemistry To Make Aggregates. Acc. Chem. Res. 1995, 28 (1), 37−44. (100) Stupp, S. I.; Pralle, M. U.; Tew, G. N.; Li, L. M.; Sayar, M.; Zubarev, E. R. Self-assembly of Organic Nano-objects into Functional Materials. MRS Bull. 2000, 25 (4), 42−48. (101) Reinhoudt, D. N.; Crego-Calama, M. Synthesis Beyond the Molecule. Science 2002, 295 (5564), 2403−2407. 11782

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Self-Assembly and Binding Properties. Chem. - Eur. J. 2005, 11 (1), 298−307. (126) Atwood, J. L.; Barbour, L. J.; Jerga, A. Hydrogen-bonded molecular capsules are stable in polar media. Chem. Commun. 2001, 2376−2377. (127) MacGillivray, L. R.; Atwood, J. L. A Chiral Spherical Molecular Assembly Held Together by 60 Hydrogen Bonds. Nature 1997, 389 (6650), 469−472. (128) Kang, J.; Santamaría, J.; Hilmersson, G.; Rebek, J. SelfAssembled Molecular Capsule Catalyzes a Diels−Alder Reaction. J. Am. Chem. Soc. 1998, 120 (29), 7389−7390. (129) Heinz, T.; Rudkevich, D. M.; Rebek, J. Pairwise Selection of Guests in a Cylindrical Molecular Capsule of Nanometre Dimensions. Nature 1998, 394 (6695), 764−766. (130) Harada, A. Cyclodextrin-Based Molecular Machines. Acc. Chem. Res. 2001, 34 (6), 456−464. (131) Mirkin, C. A.; Letsinger, R. L.; Mucic, R. C.; Storhoff, J. J. A DNA-Based Method for Rationally Assembling Nanoparticles into Macroscopic Materials. Nature 1996, 382 (6592), 607−609. (132) Zhang, X.; Rehm, S.; Safont-Sempere, M. M.; Würthner, F. Vesicular Perylene Dye Nanocapsules as Supramolecular Fluorescent pH Sensor Systems. Nat. Chem. 2009, 1 (8), 623−629. (133) Aida, T.; Meijer, E. W.; Stupp, S. I. Functional Supramolecular Polymers. Science 2012, 335 (6070), 813−817. (134) Sisson, A. L.; Shah, M. R.; Bhosale, S.; Matile, S. Synthetic Ion channels and Pores (2004−2005). Chem. Soc. Rev. 2006, 35 (12), 1269−1286. (135) Shao, L.; Zhou, J.; Hua, B.; Yu, G. A dual-responsive supraamphiphile based on a water-soluble pillar[7]arene and a naphthalene diimide-containing guest. Chem. Commun. 2015, 51 (33), 7215−7218. (136) Hill, J. P.; Jin, W.; Kosaka, A.; Fukushima, T.; Ichihara, H.; Shimomura, T.; Ito, K.; Hashizume, T.; Ishii, N.; Aida, T. SelfAssembled Hexa-peri-hexabenzocoronene Graphitic Nanotube. Science 2004, 304 (5676), 1481−1483. (137) Nolde, F.; Pisula, W.; Müller, S.; Kohl, C.; Müllen, K. Synthesis and Self-Organization of Core-Extended Perylene Tetracarboxdiimides with Branched Alkyl Substituents. Chem. Mater. 2006, 18 (16), 3715− 3725. (138) Jin, W.; Yamamoto, Y.; Fukushima, T.; Ishii, N.; Kim, J.; Kato, K.; Takata, M.; Aida, T. Systematic Studies on Structural Parameters for Nanotubular Assembly of Hexa-peri-hexabenzocoronenes. J. Am. Chem. Soc. 2008, 130 (29), 9434−9440. (139) Piot, L.; Marie, C.; Dou, X.; Feng, X.; Müllen, K.; Fichou, D. Growth of Long, Highly Stable, and Densely Packed Worm-Like Nanocolumns of Hexa-peri-Hexabenzocoronenes via Chemisorption on Au(111). J. Am. Chem. Soc. 2009, 131 (4), 1378−1379. (140) Balakrishnan, K.; Datar, A.; Oitker, R.; Chen, H.; Zuo, J.; Zang, L. Nanobelt Self-Assembly from an Organic n-Type Semiconductor: Propoxyethyl-PTCDI. J. Am. Chem. Soc. 2005, 127 (30), 10496− 10497. (141) Yan, P.; Chowdhury, A.; Holman, M. W.; Adams, D. M. SelfOrganized Perylene Diimide Nanofibers. J. Phys. Chem. B 2005, 109 (2), 724−730. (142) Zhang, X.; Chen, Z.; Würthner, F. Morphology Control of Fluorescent Nanoaggregates by Co-Self-Assembly of Wedge- and Dumbbell-Shaped Amphiphilic Perylene Bisimides. J. Am. Chem. Soc. 2007, 129 (16), 4886−4887. (143) Che, Y.; Datar, A.; Balakrishnan, K.; Zang, L. Ultralong Nanobelts Self-Assembled from an Asymmetric Perylene Tetracarboxylic Diimide. J. Am. Chem. Soc. 2007, 129 (23), 7234−7235. (144) Hoeben, F. J. M.; Zhang, J.; Lee, C. C.; Pouderoijen, M. J.; Wolffs, M.; Würthner, F.; Schenning, A. P. H. J.; Meijer, E. W.; De Feyter, S. Visualization of Various Supramolecular Assemblies of Oligo(para-phenylenevinylene)−Melamine and Perylene Bisimide. Chem. - Eur. J. 2008, 14 (28), 8579−8589. (145) Erten, Ş.; Posokhov, Y.; Alp, S.; Iċ ļ i, S. The Study of the Solubility of Naphthalene Diimides with Various Bulky flanking Substituents in Different Solvents by UV−vis Spectroscopy. Dyes Pigm. 2005, 64 (2), 171−178.

(102) Whitesides, G. M.; Grzybowski, B. Self-Assembly at All Scales. Science 2002, 295 (5564), 2418−2421. (103) Elemans, J. A. A. W.; Rowan, A. E.; Nolte, R. J. M. Mastering Molecular Matter. Supramolecular Architectures by Hierarchical Selfassembly. J. Mater. Chem. 2003, 13 (11), 2661−2670. (104) Lehn, J.-M. In Supramolecular Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 1995. (105) Pedersen, C. J. The Discovery of Crown Ethers (Noble Lecture). Angew. Chem., Int. Ed. Engl. 1988, 27 (8), 1021−1027. (106) Lehn, J.-M. Supramolecular ChemistryScope and Perspectives Molecules, Supermolecules, and Molecular Devices (Nobel Lecture). Angew. Chem., Int. Ed. Engl. 1988, 27 (1), 89−112. (107) Cram, D. J. The Design of Molecular Hosts, Guests, and Their Complexes (Nobel Lecture). Angew. Chem., Int. Ed. Engl. 1988, 27 (8), 1009−1020. (108) Tominaga, M.; Suzuki, K.; Murase, T.; Fujita, M. 24-Fold Endohedral Functionalization of a Self-Assembled M12L24 Coordination Nanoball. J. Am. Chem. Soc. 2005, 127 (34), 11950−11951. (109) Yamanoi, Y.; Sakamoto, Y.; Kusukawa, T.; Fujita, M.; Sakamoto, S.; Yamaguchi, K. Dynamic Assembly of Coordination Boxes from (en)Pd(II) Unit and a Rectangular Panel-Like Ligand: NMR, CSI-MS, and X-ray Studies. J. Am. Chem. Soc. 2001, 123 (5), 980−981. (110) Kawano, M.; Kobayashi, Y.; Ozeki, T.; Fujita, M. Direct Crystallographic Observation of a Coordinatively Unsaturated Transition-Metal Complex in situ Generated within a Self-Assembled Cage. J. Am. Chem. Soc. 2006, 128 (20), 6558−6559. (111) Takaoka, K.; Kawano, M.; Ozeki, T.; Fujita, M. Crystallographic Observation of an Olefin Photodimerization Reaction That Takes Place via Thermal Molecular Tumbling Within a Self-assembled Host. Chem. Commun. 2006, 1625−1627. (112) Ariga, K.; Hill, J. P.; Lee, M.; Vinu, A.; Charvet, R.; Acharya, S. Challenges and Breakthroughs in Recent Research on Self-Assembly. Sci. Technol. Adv. Mater. 2008, 9 (1), 014109. (113) Ajayaghosh, A.; Praveen, V. K. π-Organogels of Self-Assembled p-Phenylenevinylenes: Soft Materials with Distinct Size, Shape, and Functions. Acc. Chem. Res. 2007, 40 (8), 644−656. (114) Chakrabarty, R.; Mukherjee, P. S.; Stang, P. J. Supramolecular Coordination: Self-Assembly of Finite Two- and Three-Dimensional Ensembles. Chem. Rev. 2011, 111 (11), 6810−6918. (115) Ajayaghosh, A.; George, J. S.; Schenning, P. H. J. A. In Supermolecular Dye Chemistry, Topics in Current Chemistry; Würthner, F., Ed.; Springer: Berlin Heidelberg, Berlin, Heidelberg, 2005. (116) Carroll, R. L.; Gorman, C. B. The Genesis of Molecular Electronics. Angew. Chem., Int. Ed. 2002, 41 (23), 4378−4400. (117) Hoeben, F. J. M.; Jonkheijm, P.; Meijer, E. W.; Schenning, A. P. H. J. About Supramolecular Assemblies of π-Conjugated Systems. Chem. Rev. 2005, 105 (4), 1491−1546. (118) Schenning, A. P. H. J.; Meijer, E. W. Supramolecular Electronics; Nanowires from Self-assembled π-Conjugated Systems. Chem. Commun. 2005, 3245−3258. (119) Tour, J. M. In Molecular Electronics; World Scientific: River Edge, NJ, 2003. (120) Forrest, S. R. The Path to Ubiquitous and Low-Cost Organic Electronic Appliances on Plastic. Nature 2004, 428 (6986), 911−918. (121) Müllen, K.; Scherf, U. In Organic Light Emitting Devices; Müllen, K., Scherf, U., Eds.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2006. (122) Müllen, K.; Wegner, G. In Electronic Materials: The Oligomer Approach; Wiley-VCH Verlag GmbH: Weinheim, Germany, 1998. (123) Fujita, M.; Tominaga, M.; Hori, A.; Therrien, B. Coordination Assemblies from a Pd(II)-Cornered Square Complex. Acc. Chem. Res. 2005, 38 (4), 369−378. (124) Kamiya, N.; Tominaga, M.; Sato, S.; Fujita, M. SaccharideCoated M12L24 Molecular Spheres That Form Aggregates by Multiinteraction with Proteins. J. Am. Chem. Soc. 2007, 129 (13), 3816− 3817. (125) Corbellini, F.; Knegtel, R. M. A.; Grootenhuis, P. D. J.; CregoCalama, M.; Reinhoudt, D. N. Water-Soluble Molecular Capsules: 11783

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

(146) Zhong, C. J.; Kwan, W. S. V.; Miller, L. L. Self-assembly of delocalized π-Stacks in Solution. Assessment of Structural Effects. Chem. Mater. 1992, 4 (6), 1423−1428. (147) Shao, H.; Parquette, J. R. A π-conjugated hydrogel based on an Fmoc-dipeptide naphthalene diimide semiconductor. Chem. Commun. 2010, 46 (24), 4285−4287. (148) Shao, H.; Nguyen, T.; Romano, N. C.; Modarelli, D. A.; Parquette, J. R. Self-Assembly of 1-D n-Type Nanostructures Based on Naphthalene Diimide-Appended Dipeptides. J. Am. Chem. Soc. 2009, 131 (45), 16374−16376. (149) Shao, H.; Gao, M.; Kim, S. H.; Jaroniec, C. P.; Parquette, J. R. Aqueous Self-Assembly of L-Lysine-Based Amphiphiles into 1D nType Nanotubes. Chem. - Eur. J. 2011, 17 (46), 12882−12885. (150) Molla, M. R.; Ghosh, S. Hydrogen-Bonding-Mediated Vesicular Assembly of Functionalized Naphthalene−Diimide-Based Bolaamphiphile and Guest-Induced Gelation in Water. Chem. - Eur. J. 2012, 18 (32), 9860−9869. (151) Pandeeswar, M.; Khare, H.; Ramakumar, S.; Govindaraju, T. Crystallographic Insight-Guided Nanoarchitectonics and Conductivity Modulation of an n-Type Organic Semiconductor Through Peptide Conjugation. Chem. Commun. 2015, 51 (39), 8315−8318. (152) Shao, H.; Seifert, J.; Romano, N. C.; Gao, M.; Helmus, J. J.; Jaroniec, C. P.; Modarelli, D. A.; Parquette, J. R. Amphiphilic SelfAssembly of an n-Type Nanotube. Angew. Chem., Int. Ed. 2010, 49 (42), 7688−7691. (153) Kim, S. H.; Parquette, J. R. 1D Self-assembly of Terthiophene (3T)-Naphthalenediimide (NDI) Dyad. Chem. Lett. 2014, 43 (10), 1634−1636. (154) Bhosale, S. V.; Al Kobaisi, M.; Bhosale, R. S.; Bhosale, S. V. Well-Organized Supramolecular Self-assembly of Acene Diimide Derivatives. RSC Adv. 2013, 3 (43), 19840−19843. (155) Nandre, K. P.; Al Kobaisi, M.; Bhosale, R. S.; Latham, K.; Bhosale, S. V.; Bhosale, S. V. pH Triggered Self-assembly Induced Enhanced Emission of Phosphonic Acid Appended Naphthalenediimide Amphiphile. RSC Adv. 2014, 4 (76), 40381−40384. (156) Avinash, M. B.; Govindaraju, T. Molecular Engineering: Engineering Molecular Organization of Naphthalenediimides: Large Nanosheets with Metallic Conductivity and Attoliter Containers. Adv. Funct. Mater. 2011, 21 (20), 3875−3882. (157) Avinash, M. B.; Govindaraju, T. A Bio-inspired Design Strategy: Organization of Tryptophan-Appended Naphthalenediimide into Well-defined Architectures Induced by Molecular Interactions. Nanoscale 2011, 3 (6), 2536−2543. (158) Wang, F.-X.; Liu, Y.-Q.; Qiu, S.; Pan, G.-B. Solvent Effect on Hierarchical Assembly of 2-Aminooctane-Functionalized Naphthalenediimide. RSC Adv. 2014, 4 (12), 6009−6013. (159) Pandeeswar, M.; Avinash, M. B.; Govindaraju, T. Chiral Transcription and Retentive Helical Memory: Probing Peptide Auxiliaries Appended with Naphthalenediimides for Their OneDimensional Molecular Organization. Chem. - Eur. J. 2012, 18 (16), 4818−4822. (160) Mondal, T.; Basak, D.; Al Ouahabi, A.; Schmutz, M.; Mesini, P.; Ghosh, S. Extended Supramolecular Organization of π-Systems Using yet Unexplored Simultaneous Intra- and Inter-Molecular HBonding Motifs of 1,3-Dihydroxy Derivatives. Chem. Commun. 2015, 51 (24), 5040−5043. (161) Lee, K. S.; Parquette, J. R. A Self-Assembled Nanotube for the Direct Aldol Reaction in Water. Chem. Commun. 2015, 51 (86), 15653−15656. (162) Avinash, M. B.; Swathi, K.; Narayan, K. S.; Govindaraju, T. Molecular Architectonics of Naphthalenediimides for Efficient Structure−Property Correlation. ACS Appl. Mater. Interfaces 2016, 8 (13), 8678−8685. (163) Salazar-Rios, J. M.; Gomulya, W.; Derenskyi, V.; Yang, J.; Bisri, S. Z.; Chen, Z.; Facchetti, A.; Loi, M. A. Selecting Semiconducting Single-Walled Carbon Nanotubes with Narrow Bandgap Naphthalene Diimide-Based Polymers. Adv. Electron. Mater. 2015, 1 (8), 1500074. (164) Hu, Z.; Pantoş, G. D.; Kuganathan, N.; Arrowsmith, R. L.; Jacobs, R. M. J.; Kociok-Köhn, G.; O’Byrne, J.; Jurkschat, K.; Burgos,

P.; Tyrrell, R. M.; Botchway, S. W.; Sanders, J. K. M.; Pascu, S. I. Interactions Between Amino Acid-Tagged Naphthalenediimide and Single Walled Carbon Nanotubes for the Design and Construction of New Bioimaging Probes. Adv. Funct. Mater. 2012, 22 (3), 503−518. (165) Castaldelli, E.; Triboni, E. R.; Demets, G. J.-F. Self-Assembled Naphthalenediimide Derivative Films for Light-Assisted Electrochemical Reduction of Oxygen. Chem. Commun. 2011, 47 (19), 5581−5583. (166) Billeci, F.; D’Anna, F.; Marullo, S.; Noto, R. Self-assembly of Fluorescent Diimidazolium Salts: Tailor Properties of the Aggregates Changing Alkyl Chain Features. RSC Adv. 2016, 6, 59502. (167) Hsu, L.-H.; Hsu, S.-M.; Wu, F.-Y.; Liu, Y.-H.; Nelli, S. R.; Yeh, M.-Y.; Lin, H.-C. Nanofibrous Hydrogels Self-assembled from Naphthalene Diimide (NDI)/Amino Acid Conjugates. RSC Adv. 2015, 5 (26), 20410−20413. (168) Shankar, B. H.; Jayaram, D. T.; Ramaiah, D. Naphthalene Imide Conjugates: Formation of Supramolecular Assemblies, and the Encapsulation and Release of Dyes through DNA-Mediated Disassembly. Chem. - Eur. J. 2015, 21 (49), 17657−17663. (169) Pfeffermann, M.; Dong, R.; Graf, R.; Zajaczkowski, W.; Gorelik, T.; Pisula, W.; Narita, A.; Müllen, K.; Feng, X. Free-Standing Monolayer Two-Dimensional Supramolecular Organic Framework with Good Internal Order. J. Am. Chem. Soc. 2015, 137 (45), 14525− 14532. (170) Basak, S.; Nanda, J.; Banerjee, A. Assembly of Naphthalenediimide Conjugated Peptides: Aggregation Induced Changes in Fluorescence. Chem. Commun. 2013, 49 (61), 6891−6893. (171) Basak, S.; Bhattacharya, S.; Datta, A.; Banerjee, A. ChargeTransfer Complex Formation in Gelation: The Role of Solvent Molecules with Different Electron-Donating Capacities. Chem. - Eur. J. 2014, 20 (19), 5721−5726. (172) Basak, S.; Nandi, N.; Baral, A.; Banerjee, A. Tailor-Made Design of J- or H-Aggregated Naphthalenediimide-Based Gels and Remarkable Fluorescence Turn on/off Behaviour Depending on Solvents. Chem. Commun. 2015, 51 (4), 780−783. (173) Liu, Y.-H.; Hsu, S.-M.; Wu, F.-Y.; Cheng, H.; Yeh, M.-Y.; Lin, H.-C. Electroactive Organic Dye Incorporating Dipeptides in the Formation of Self-Assembled Nanofibrous Hydrogels. Bioconjugate Chem. 2014, 25 (10), 1794−1800. (174) Kumar, M.; George, S. J. Spectroscopic Probing of the Dynamic Self-Assembly of an Amphiphilic Naphthalene Diimide Exhibiting Reversible Vapochromism. Chem. - Eur. J. 2011, 17 (40), 11102−11106. (175) Kumar, M.; George, S. J. Green Fluorescent Organic Nanoparticles by Self-Assembly Induced Enhanced Emission of a Naphthalene Diimide Bolaamphiphile. Nanoscale 2011, 3 (5), 2130− 2133. (176) Kulkarni, C.; George, S. J. Carbonate Linkage Bearing Naphthalenediimides: Self-Assembly and Photophysical Properties. Chem. - Eur. J. 2014, 20 (16), 4537−4541. (177) Kulkarni, C.; Periyasamy, G.; Balasubramanian, S.; George, S. J. Charge-Transfer Complexation between Naphthalene Diimides and Aromatic Solvents. Phys. Chem. Chem. Phys. 2014, 16 (28), 14661− 14664. (178) Kumar, M.; George, S. J. Novel Coronene−Naphthalene Dimide-Based Donor−Acceptor Pair for Tunable Charge-Transfer Nanostructures. Chem. - Asian J. 2014, 9 (9), 2427−2431. (179) Molla, M. R.; Ghosh, S. Hydrogen-Bonding-Mediated JAggregation and White-Light Emission from a Remarkably Simple, Single-Component, Naphthalenediimide Chromophore. Chem. - Eur. J. 2012, 18 (5), 1290−1294. (180) Molla, M. R.; Gehrig, D.; Roy, L.; Kamm, V.; Paul, A.; Laquai, F.; Ghosh, S. Self-Assembly of Carboxylic Acid Appended Naphthalene Diimide Derivatives with Tunable Luminescent Color and Electrical Conductivity. Chem. - Eur. J. 2014, 20 (3), 760−771. (181) Molla, M. R.; Ghosh, S. Structural Variations on Self-Assembly and Macroscopic Properties of 1,4,5,8-Naphthalene-diimide Chromophores. Chem. Mater. 2011, 23 (1), 95−105. 11784

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

(201) Das, A.; Ghosh, S. Stimuli-Responsive Self-Assembly of a Naphthalene Diimide by Orthogonal Hydrogen Bonding and Its Coassembly with a Pyrene Derivative by a Pseudo-Intramolecular Charge-Transfer Interaction. Angew. Chem. 2014, 126 (4), 1110− 1115. (202) Molla, M. R.; Das, A.; Ghosh, S. Chiral Induction by Helical Neighbour: Spectroscopic Visualization of Macroscopic-Interaction Among Self-Sorted Donor and Acceptor π-Stacks. Chem. Commun. 2011, 47 (31), 8934−8936. (203) Kar, H.; Ghosh, S. Three Component Assemblies by Orthogonal H-Bonding and Donor-Acceptor Charge-Transfer Interaction. Chem. Commun. 2014, 50 (9), 1064−1066. (204) Das, A.; Molla, M. R.; Maity, B.; Koley, D.; Ghosh, S. Hydrogen-Bonding Induced Alternate Stacking of Donor (D) and Acceptor (A) Chromophores and their Supramolecular Switching to Segregated States. Chem. - Eur. J. 2012, 18 (32), 9849−9859. (205) Liu, K.; Wang, C.; Li, Z.; Zhang, X. Superamphiphiles Based on Directional Charge-Transfer Interactions: From Supramolecular Engineering to Well-Defined Nanostructures. Angew. Chem., Int. Ed. 2011, 50 (21), 4952−4956. (206) Liu, K.; Yao, Y.; Liu, Y.; Wang, C.; Li, Z.; Zhang, X. SelfAssembly of Supra-amphiphiles Based on Dual Charge-Transfer Interactions: From Nanosheets to Nanofibers. Langmuir 2012, 28 (29), 10697−10702. (207) Liu, K.; Yao, Y.; Wang, C.; Liu, Y.; Li, Z.; Zhang, X. From Bolaamphiphiles to Supra-amphiphiles: The Transformation from TwoDimensional Nanosheets into One-Dimensional Nanofibers with Tunable-Packing Fashion of n-Type Chromophores. Chem. - Eur. J. 2012, 18 (28), 8622−8628. (208) Kumar, M.; Ushie, O. A.; George, S. J. Supramolecular Clippers for Controlling Photophysical Processes through Preorganized Chromophores. Chem. - Eur. J. 2014, 20 (17), 5141−5148. (209) Yu, G.; Li, J.; Yu, W.; Han, C.; Mao, Z.; Gao, C.; Huang, F. Carbon Nanotube/Biocompatible Bola-Amphiphile Supramolecular Biohybrid Materials: Preparation and Their Application in Bacterial Cell Agglutination. Adv. Mater. 2013, 25 (44), 6373−6379. (210) Koshkakaryan, G.; Klivansky, L. M.; Cao, D.; Snauko, M.; Teat, S. J.; Struppe, J. O.; Liu, Y. Alternative Donor−Acceptor Stacks from Crown Ethers and Naphthalene Diimide Derivatives: Rapid, Selective Formation from Solution and Solid State Grinding. J. Am. Chem. Soc. 2009, 131 (6), 2078−2079. (211) Fox, J.; Wie, J. J.; Greenland, B. W.; Burattini, S.; Hayes, W.; Colquhoun, H. M.; Mackay, M. E.; Rowan, S. J. High-Strength, Healable, Supramolecular Polymer Nanocomposites. J. Am. Chem. Soc. 2012, 134 (11), 5362−5368. (212) Avinash, M. B.; Sandeepa, K. V.; Govindaraju, T. Molecular Assembly of Amino Acid Interlinked, Topologically Symmetric, πComplementary Donor−Acceptor−Donor Triads. Beilstein J. Org. Chem. 2013, 9, 1565−1571. (213) Prentice, G. M.; Pascu, S. I.; Filip, S. V.; West, K. R.; Pantos, G. D. Aromatic Donor-Acceptor Interactions in Non-Polar Environments. Chem. Commun. 2015, 51 (39), 8265−8268. (214) Sanders, A. M.; Magnanelli, T. J.; Bragg, A. E.; Tovar, J. D. Photoinduced Electron Transfer within Supramolecular Donor− Acceptor Peptide Nanostructures under Aqueous Conditions. J. Am. Chem. Soc. 2016, 138 (10), 3362−3370. (215) Yamada, J.-i.; Akutsu, H.; Nishikawa, H.; Kikuchi, K. New Trends in the Synthesis of π-Electron Donors for Molecular Conductors and Superconductors. Chem. Rev. 2004, 104 (11), 5057−5084. (216) Dirian, K.; Backes, S.; Backes, C.; Strauss, V.; Rodler, F.; Hauke, F.; Hirsch, A.; Guldi, D. M. Naphthalenebisimides as Photofunctional Surfactants for SWCNTs - Towards Water-Soluble Electron Donor-Acceptor Hybrids. Chem. Sci. 2015, 6 (12), 6886− 6895. (217) Schmidt, B. M.; Osuga, T.; Sawada, T.; Hoshino, M.; Fujita, M. Compressed Corannulene in a Molecular Cage. Angew. Chem., Int. Ed. 2016, 55 (4), 1561−1564.

(182) Kar, H.; Ghosh, S. J-aggregation of a Sulfur-Substituted Naphthalenediimide (NDI) with Remarkably Bright Fluorescence. Chem. Commun. 2016, 52 (57), 8818−8821. (183) Sakurai, T.; Tsutsui, Y.; Kato, K.; Takata, M.; Seki, S. Preferential Formation of Columnar Mesophases via Peripheral Modification of Discotic π-Systems with Immiscible Side Chain Pairs. J. Mater. Chem. C 2016, 4 (7), 1490−1496. (184) Das, A.; Molla, M.; Ghosh, S. Comparative Self-Assembly Studies and Self-Sorting of two Structurally Isomeric NaphthaleneDiimide (NDI)-Gelators. J. Chem. Sci. 2011, 123 (6), 963−973. (185) Das, A.; Ghosh, S. Luminescent Invertible Polymersome by Remarkably Stable Supramolecular Assembly of Naphthalene Diimide (NDI) π-System. Macromolecules 2013, 46 (10), 3939−3949. (186) Rajdev, P.; Molla, M. R.; Ghosh, S. Understanding the Role of H-Bonding in Aqueous Self-Assembly of Two Naphthalene Diimide (NDI)-Conjugated Amphiphiles. Langmuir 2014, 30 (8), 1969−1976. (187) Das, A.; Ghosh, S. To Assemble or Fold? Chem. Commun. 2014, 50 (79), 11657−11660. (188) Shinde, S. V.; Kulkarni, M.; Talukdar, P. Helical Supramolecular Organization of a 1,2-Diol Appended Naphthalene Diimide Organogelator via an Extended Intermolecular H-Bonding Network. RSC Adv. 2016, 6 (36), 30690−30694. (189) Bhosale, S. V.; Jani, C.; Lalander, C. H.; Langford, S. J. Solvophobic Control of Core-Substituted Naphthalene Diimide Nanostructures. Chem. Commun. 2010, 46 (6), 973−975. (190) Bhosale, S. V.; Jani, C. H.; Lalander, C. H.; Langford, S. J.; Nerush, I.; Shapter, J. G.; Villamaina, D.; Vauthey, E. Supramolecular Construction of Vesicles Based on Core-Substituted Naphthalene Diimide Appended with Triethyleneglycol Motifs. Chem. Commun. 2011, 47 (29), 8226−8228. (191) De Cat, I.; Roger, C.; Lee, C. C.; Hoeben, F. J. M.; Pouderoijen, M. J.; Schenning, A. P. H. J.; Würthner, F.; De Feyter, S. Identification of Oligo(p-Phenylene vinylene)-Naphthalene Diimide Heterocomplexes by Scanning Tunneling Microscopy and Spectroscopy at the Liquid-Solid Interface. Chem. Commun. 2008, 5496−5498. (192) Xu, J.; Yu, H.; Yang, L.; Wu, G.; Wang, Z.; Wang, D.; Zhang, X. Self-assembling 1D Core/Shell Microrods by the Introduction of Additives: a One-pot and Shell-tunable Method. Chem. Sci. 2015, 6 (8), 4907−4911. (193) Wu, Y.-C.; Leowanawat, P.; Sun, H.-J.; Partridge, B. E.; Peterca, M.; Graf, R.; Spiess, H. W.; Zeng, X.; Ungar, G.; Hsu, C.-S.; Heiney, P. A.; Percec, V. Complex Columnar Hexagonal Polymorphism in Supramolecular Assemblies of a Semifluorinated Electron-Accepting Naphthalene Bisimide. J. Am. Chem. Soc. 2015, 137 (2), 807−819. (194) Gu, P.-Y.; Wang, Z.; Liu, G.; Nie, L.; Ganguly, R.; Li, Y.; Zhang, Q. Synthesis, Physical Properties, and Sensing Behaviour of a Novel Naphthalenediimide Derivative. Dyes Pigm. 2016, 131, 224− 230. (195) Kar, H.; Gehrig, D. W.; Laquai, F.; Ghosh, S. J-Aggregation, its Impact on Excited State Dynamics and Unique Solvent Effects on Macroscopic Assembly of a Core-Substituted Naphthalenediimide. Nanoscale 2015, 7 (15), 6729−6736. (196) Kar, H.; Gehrig, D. W.; Allampally, N. K.; Fernandez, G.; Laquai, F.; Ghosh, S. Cooperative Supramolecular Polymerization of an Amine-Substituted Naphthalene-Diimide and its Impact on Excited State Photophysical Properties. Chem. Sci. 2016, 7 (2), 1115−1120. (197) Kumar, N. S. S.; Gujrati, M. D.; Wilson, J. N. Evidence of Preferential π-Stacking: a Study of Intermolecular and Intramolecular Charge Transfer Complexes. Chem. Commun. 2010, 46 (30), 5464− 5466. (198) Reczek, J. J.; Iverson, B. L. Using Aromatic Donor Acceptor Interactions to Affect Macromolecular Assembly. Macromolecules 2006, 39 (17), 5601−5603. (199) Gabriel, G. J.; Sorey, S.; Iverson, B. L. Altering the Folding Patterns of Naphthyl Trimers. J. Am. Chem. Soc. 2005, 127 (8), 2637− 2640. (200) Das, A.; Ghosh, S. A Generalized Supramolecular Strategy for Self-Sorted Assembly Between Donor and Acceptor Gelators. Chem. Commun. 2011, 47 (31), 8922−8924. 11785

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

(218) Zhang, L.; Zhang, Z.; He, C.; Dai, L.; Liu, J.; Wang, L. Rationally Designed Surfactants for Few-Layered Graphene Exfoliation: Ionic Groups Attached to Electron-Deficient π-Conjugated Unit through Alkyl Spacers. ACS Nano 2014, 8 (7), 6663−6670. (219) Chen, M.; Yang, C.; Xu, Z.; Tang, Y.; Jiang, J.; Liu, P.; Su, Y.; Wu, D. A Facile Self-Assembly Strategy Towards Naphthalene Diimide/Graphene Hybrids as High Performance Organic Cathodes for Lithium-Ion Batteries. RSC Adv. 2016, 6 (17), 13666−13669. (220) Bhattacharjee, S.; Bhattacharya, S. Charge Transfer Induces Formation of Stimuli-Responsive, Chiral, Cohesive Vesicles-on-aString that Eventually Turn into a Hydrogel. Chem. - Asian J. 2015, 10 (3), 572−580. (221) Bhattacharjee, S.; Maiti, B.; Bhattacharya, S. First Report of Charge-Transfer Induced Heat-Set Hydrogel. Structural Insights and Remarkable Properties. Nanoscale 2016, 8 (21), 11224−11233. (222) Sikder, A.; Ghosh, B.; Chakraborty, S.; Paul, A.; Ghosh, S. Rational Design for Complementary Donor−Acceptor Recognition Pairs using Self-Complementary Hydrogen Bonds. Chem. - Eur. J. 2016, 22, 1908−1913. (223) Sikdar, N.; Jayaramulu, K.; Kiran, V.; Rao, K. V.; Sampath, S.; George, S. J.; Maji, T. K. Redox-Active Metal−Organic Frameworks: Highly Stable Charge-Separated States through Strut/Guest-to-Strut Electron Transfer. Chem. - Eur. J. 2015, 21 (33), 11701−11706. (224) Fan, J.; Chang, X.; He, M.; Shang, C.; Wang, G.; Yin, S.; Peng, H.; Fang, Y. Functionality Oriented Derivatization of Naphthalenediimide: A Molecular Gel Strategy-Based Fluorescent Film for Aniline Vapor Detection. ACS Appl. Mater. Interfaces 2016, 8, 18584. (225) Lee, J. H.; Park, J.; Jung, J. H. Control of Luminescence Properties in Naphthalene Diimide-Based Gel with Azodibenzoic Acid by Charge Transfer Interaction. Bull. Korean Chem. Soc. 2014, 35 (9), 2851−2854. (226) Pan, M.; Lin, X.-M.; Li, G.-B.; Su, C.-Y. Progress in the Study of Metal−Organic Materials Applying Naphthalene Diimide (NDI) Ligands. Coord. Chem. Rev. 2011, 255 (15−16), 1921−1936. (227) Ulrich, S.; Petitjean, A.; Lehn, J.-M. Metallo-Controlled Dynamic Molecular Tweezers: Design, Synthesis, and Self-Assembly by Metal-Ion Coordination. Eur. J. Inorg. Chem. 2010, 2010 (13), 1913−1928. (228) Deng, H.-Y.; He, J.-R.; Pan, M.; Li, L.; Su, C.-Y. Synergistic Metal and Anion Effects on the Formation of Coordination Assemblies from a N,N′-bis(3-Pyridylmethyl)naphthalene Diimide Ligand. CrystEngComm 2009, 11 (5), 909−917. (229) Li, G.-B.; He, J.-R.; Pan, M.; Deng, H.-Y.; Liu, J.-M.; Su, C.-Y. Construction of 0D to 3D Cadmium Complexes from Different Pyridyl Diimide Ligands. Dalton Trans. 2012, 41 (15), 4626−4633. (230) Li, G.-B.; Liu, J.-M.; Cai, Y.-P.; Su, C.-Y. Structural Diversity of a Series of Mn(II), Cd(II), and Co(II) Complexes with Pyridine Donor Diimide Ligands. Cryst. Growth Des. 2011, 11 (7), 2763−2772. (231) Han, L.; Qin, L.; Xu, L.; Zhou, Y.; Sun, J.; Zou, X. A Novel Photochromic Calcium-Based Metal-Organic Framework Derived from a Naphthalene Diimide Chromophore. Chem. Commun. 2013, 49 (4), 406−408. (232) Yu, Y.; Li, Y.; Chen, S.; Liu, T.; Qin, Z.; Liu, H.; Li, Y. Synthesis of a Naphthalene-diimide Cyclophane for Tuning Supramolecular Interactions by Metal Ions. Eur. J. Org. Chem. 2012, 2012 (23), 4287−4292. (233) Keller, J. M.; Schanze, K. S. Synthesis of Monodisperse Platinum Acetylide Oligomers End-Capped with Naphthalene Diimide Units. Organometallics 2009, 28 (14), 4210−4216. (234) Liao, C.; Yarnell, J. E.; Glusac, K. D.; Schanze, K. S. Photoinduced Charge Separation in Platinum Acetylide Oligomers. J. Phys. Chem. B 2010, 114 (45), 14763−14771. (235) Lee, K. A.; Lozan, V.; Langford, S. J.; Kersting, B. Ternary Complexes Composed of Naphthalene Diimides and Binucleating Metallocavitands: Preparation, Characterisation and Structure of [(Ni2L)2(NDI)][BPh4]2. Dalton Trans. 2009, No. 36, 7481−7485. (236) Shokouhi Mehr, H.; Romano, N. C.; Altamimi, R.; Modarelli, J. M.; Modarelli, D. A. Core Substituted Naphthalene Diimide - Metallo

Bisterpyridine Supramolecular Polymers: Synthesis, Photophysics and Morphology. Dalton Trans. 2015, 44 (7), 3176−3184. (237) Liu, J.-J.; Guan, Y.-F.; Chen, Y.; Lin, M.-J.; Huang, C.-C.; Dai, W.-X. The Impact of Lone Pair-π Interactions on Photochromic Properties in 1-D Naphthalene Diimide Coordination Networks. Dalton Trans. 2015, 44 (39), 17312−17317. (238) Bauer, E. M.; Bellitto, C.; Gómez García, C. J.; Righini, G. Synthesis and Characterization of a New Layered Organic−Inorganic Hybrid Nickel(II) 1,4:5,8-Naphthalenediimide bis-Phosphonate, Exhibiting Canted Antiferromagnetism, with Tc ∼ 21 K. J. Solid State Chem. 2008, 181 (5), 1213−1219. (239) Cairns, A. J.; Perman, J. A.; Wojtas, L.; Kravtsov, V. C.; Alkordi, M. H.; Eddaoudi, M.; Zaworotko, M. J. Supermolecular Building Blocks (SBBs) and Crystal Design: 12-Connected Open Frameworks Based on a Molecular Cubohemioctahedron. J. Am. Chem. Soc. 2008, 130 (5), 1560−1561. (240) Mallick, A.; Garai, B.; Addicoat, M. A.; Petkov, P. S.; Heine, T.; Banerjee, R. Solid State Organic Amine Detection in a Photochromic Porous Metal Organic Framework. Chem. Sci. 2015, 6 (2), 1420− 1425. (241) Garai, B.; Mallick, A.; Banerjee, R. Photochromic MetalOrganic Frameworks for Inkless and Erasable Printing. Chem. Sci. 2016, 7 (3), 2195−2200. (242) Perman, J. A.; Cairns, A. J.; Wojtas, L.; Eddaoudi, M.; Zaworotko, M. J. Cocrystal Controlled Solid-State Synthesis of a Rigid Tetracarboxylate Ligand That Pillars Both Square Grid and Kagome Lattice Layers. CrystEngComm 2011, 13 (9), 3130−3133. (243) Xie, Y.-X.; Zhao, W.-N.; Li, G.-C.; Liu, P.-F.; Han, L. A Naphthalenediimide-Based Metal−Organic Framework and thin-film Exhibiting Photochromic and Electrochromic Properties. Inorg. Chem. 2016, 55 (2), 549−551. (244) Liu, J.-J.; Chen, Y.; Lin, M.-J.; Huang, C.-C.; Dai, W.-X. TwoSemiconductive-Component Hybrid Coordination Polymers with Controllable Photo-Induced Electron-Transfer Properties. Dalton Trans. 2016, 45 (15), 6339−6342. (245) Guo, Z.; Panda, D. K.; Maity, K.; Lindsey, D.; Parker, T. G.; Albrecht-Schmitt, T. E.; Barreda-Esparza, J. L.; Xiong, P.; Zhou, W.; Saha, S. Modulating the Electrical Conductivity of Metal-Organic Framework Films with Intercalated Guest π-Systems. J. Mater. Chem. C 2016, 4 (5), 894−899. (246) Palma, C.-A.; Bjork, J.; Bonini, M.; Dyer, M. S.; Llanes-Pallas, A.; Bonifazi, D.; Persson, M.; Samorì, P. Tailoring Bicomponent Supramolecular Nanoporous Networks: Phase Segregation, Polymorphism, and Glasses at the Solid−Liquid Interface. J. Am. Chem. Soc. 2009, 131 (36), 13062−13071. (247) Ruiz-Osés, M.; de Oteyza, D. G.; Fernández-Torrente, I.; Gonzalez-Lakunza, N.; Schmidt-Weber, P. M.; Kampen, T.; Horn, K.; Gourdon, A.; Arnau, A.; Ortega, J. E. Non-Covalent Interactions in Supramolecular Assemblies Investigated with Electron Spectroscopies. ChemPhysChem 2009, 10 (6), 896−900. (248) Bulheller, B. M.; Pantos, G. D.; Sanders, J. K. M.; Hirst, J. D. Electronic Structure and Circular Dichroism Spectroscopy of Naphthalenediimide Nanotubes. Phys. Chem. Chem. Phys. 2009, 11 (29), 6060−6065. (249) Anderson, T. W.; Sanders, J. K. M.; Pantos, G. D. The Sergeants-and-Soldiers Effect: Chiral Amplification in Naphthalenediimide Nanotubes. Org. Biomol. Chem. 2010, 8 (19), 4274−4280. (250) Anderson, T. W.; Pantos, G. D.; Sanders, J. K. M. Supramolecular Chemistry of Monochiral Naphthalenediimides. Org. Biomol. Chem. 2011, 9 (21), 7547−7553. (251) Stefankiewicz, A. R.; Tamanini, E.; Pantoş, G. D.; Sanders, J. K. M. Proton-Driven Switching Between Receptors for C60 and C70. Angew. Chem., Int. Ed. 2011, 50 (25), 5725−5728. (252) Kumar, M.; Jonnalagadda, N.; George, S. J. Molecular Recognition Driven Self-Assembly and Chiral Induction in Naphthalene Diimide Amphiphiles. Chem. Commun. 2012, 48 (89), 10948− 10950. (253) Nandre, K. P.; Bhosale, S. V.; Rama Krishna, K. V. S.; Gupta, A.; Bhosale, S. V. A Phosphonic Acid Appended Naphthalene Diimide 11786

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Motif for Self-Assembly into Tunable Nanostructures Through Molecular Recognition with Arginine in Water. Chem. Commun. 2013, 49 (48), 5444−5446. (254) Bhosale, R. S.; Al Kobaisi, M.; Bhosale, S. V.; Bhargava, S.; Bhosale, S. V. Flower-Like Supramolecular Self-assembly of Phosphonic Acid Appended Naphthalene Diimide and Melamine. Sci. Rep. 2015, 5, 14609. (255) Watanabe, S.; Sato, S.; Ohtsuka, K.; Takenaka, S. Electrochemical DNA Analysis with a Supramolecular Assembly of Naphthalene Diimide, Ferrocene, and β-Cyclodextrin. Anal. Chem. 2011, 83 (19), 7290−7296. (256) Nakamura, M.; Okaue, T.; Takada, T.; Yamana, K. DNATemplated Assembly of Naphthalenediimide Arrays. Chem. - Eur. J. 2012, 18 (1), 196−201. (257) Sato, S.; Takenaka, S. Electrochemical DNA Detection Using Supramolecular Interactions. Anal. Sci. 2012, 28 (7), 643−649. (258) Bhosale, R.; Kishore, R. S. K.; Ravikumar, V.; Kel, O.; Vauthey, E.; Sakai, N.; Matile, S. Zipper Assembly of SHJ Photosystems: Focus on red Naphthalenediimides, Optoelectronic Finetuning and Topological Matching. Chem. Sci. 2010, 1 (3), 357−368. (259) Sakai, N.; Kishore, R. S. K.; Matile, S. Three-Component Zipper Assembly of Photoactive Cascade Architectures with Blue, Red and Colorless Naphthalenediimide Donors and Acceptors. Org. Biomol. Chem. 2008, 6 (21), 3970−3976. (260) Lista, M.; Sakai, N.; Matile, S. Mini-Zippers: Determination of Oligomer Effects for the Assembly of Photoactive Supramolecular Rod/Stack Architectures on Gold Nanoparticles and Gold Electrodes. Supramol. Chem. 2009, 21 (3−4), 238−244. (261) Rananaware, A.; La, D. D.; Bhosale, S. V. Solvophobic Control Aggregation-Induced Emission of Tetraphenylethene-Substituted Naphthalene Diimide via Intramolecular Charge Transfer. RSC Adv. 2015, 5 (77), 63130−63134. (262) Rananaware, A.; La, D. D.; Jackson, S. M.; Bhosale, S. V. Construction of a Highly Efficient Near-IR Solid Emitter Based on Naphthalene Diimide with AIE-Active Tetraphenylethene Periphery. RSC Adv. 2016, 6 (20), 16250−16255. (263) Ghule, N. V.; La, D. D.; Bhosale, R. S.; Al Kobaisi, M.; Raynor, A. M.; Bhosale, S. V.; Bhosale, S. V. Effect of Amide Hydrogen Bonding Interaction on Supramolecular Self-Assembly of Naphthalene Diimide Amphiphiles with Aggregation Induced Emission. ChemistryOpen 2016, 5, 157. (264) Bolag, A.; Sakai, N.; Matile, S. Dipolar Photosystems: Engineering Oriented Push−Pull Components into Double- and Triple-Channel Surface Architectures. Chem. - Eur. J. 2016, 22, 9006. (265) Aleshinloye, A. O.; Bodapati, J. B.; Icil, H. Synthesis, Characterization, Optical and Electrochemical Properties of a new Chiral Multichromophoric System Based on Perylene and Naphthalene Diimides. J. Photochem. Photobiol., A 2015, 300, 27−37. (266) Narayan, B.; Bejagam, K. K.; Balasubramanian, S.; George, S. J. Autoresolution of Segregated and Mixed p-n Stacks by Stereoselective Supramolecular Polymerization in Solution. Angew. Chem., Int. Ed. 2015, 54 (44), 13053−13057. (267) Bé, A. G.; Tran, C.; Sechrist, R.; Reczek, J. J. Strongly Dichroic Organic Films via Controlled Assembly of Modular Aromatic ChargeTransfer Liquid Crystals. Org. Lett. 2015, 17 (19), 4834−4837. (268) Pu, L. Fluorescence of Organic Molecules in Chiral Recognition. Chem. Rev. 2004, 104 (3), 1687−1716. (269) Lippert, A. R.; Van de Bittner, G. C.; Chang, C. J. Boronate Oxidation as a Bioorthogonal Reaction Approach for Studying the Chemistry of Hydrogen Peroxide in Living Systems. Acc. Chem. Res. 2011, 44 (9), 793−804. (270) Quang, D. T.; Kim, J. S. Fluoro- and Chromogenic Chemodosimeters for Heavy Metal Ion Detection in Solution and Biospecimens. Chem. Rev. 2010, 110 (10), 6280−6301. (271) Bobe, S. R.; Bhosale, R. S.; Goskulwad, S. P.; Puyad, A. L.; Bhosale, S. V.; Bhosale, S. V. A Highly Selective Colorimetric Cys Sensor Based on Core-Substituted Naphthalene Diimides. RSC Adv. 2015, 5 (122), 100697−100701.

(272) Asthana, D.; Shukla, J.; Dana, S.; Rani, V.; Ajayakumar, M. R.; Rawat, K.; Mandal, K.; Yadav, P.; Ghosh, S.; Mukhopadhyay, P. Assorted Morphosynthesis: Access to Multi-Faceted Nano-Architectures from a Super-Responsive Dual π-Functional Amphiphilic Construct. Chem. Commun. 2015, 51 (83), 15237−15240. (273) Takenaka, S.; Yamashita, K.; Takagi, M.; Uto, Y.; Kondo, H. DNA Sensing on a DNA Probe-Modified Electrode Using Ferrocenylnaphthalene Diimide as the Electrochemically Active Ligand. Anal. Chem. 2000, 72 (6), 1334−1341. (274) You, M.-H.; Yuan, X.; Fang, X.; Lin, M.-J. Lone Pair-π Interactions in Naphthalene Diimide π-Acid Dyes. Supramol. Chem. 2015, 27 (7−8), 460−464. (275) Giese, M.; Albrecht, M.; Rissanen, K. Anion−π Interactions with Fluoroarenes. Chem. Rev. 2015, 115 (16), 8867−8895. (276) Estarellas, C.; Frontera, A.; Quiñonero, D.; Deyà, P. M. Relevant Anion−π Interactions in Biological Systems: The Case of Urate Oxidase. Angew. Chem., Int. Ed. 2011, 50 (2), 415−418. (277) Giese, M.; Albrecht, M.; Rissanen, K. Experimental Investigation of Anion-π Interactions - Applications and Biochemical Relevance. Chem. Commun. 2016, 52 (9), 1778−1795. (278) Jana, P.; Maity, S. K.; Bera, S.; Ghorai, P. K.; Haldar, D. Hierarchical Self-Assembly of Naphthalene Bisimides to Fluorescent Microspheres and Fluoride Sensing. CrystEngComm 2013, 15 (13), 2512−2518. (279) Bhosale, S. V.; Bhosale, S. V.; Kalyankar, M. B.; Langford, S. J. A Core-Substituted Naphthalene Diimide Fluoride Sensor. Org. Lett. 2009, 11 (23), 5418−5421. (280) Buckland, D.; Bhosale, S. V.; Langford, S. J. A Chemodosimer Based on a Core-substituted Naphthalene Diimide for Fluoride Ion Detection. Tetrahedron Lett. 2011, 52 (16), 1990−1992. (281) Bobe, S. R.; Bhosale, S. V.; Jones, L.; Puyad, A. L.; Raynor, A. M.; Bhosale, S. V. A Near-Infrared Fluoride Sensor Based on a Substituted Naphthalenediimide−Anthraquinone Conjugate. Tetrahedron Lett. 2015, 56 (33), 4762−4766. (282) Yoo, J.; Kim, M.-S.; Hong, S.-J.; Sessler, J. L.; Lee, C.-H. Selective Sensing of Anions with Calix[4]pyrroles Strapped with Chromogenic Dipyrrolylquinoxalines. J. Org. Chem. 2009, 74 (3), 1065−1069. (283) Gorteau, V.; Bollot, G.; Mareda, J.; Perez-Velasco, A.; Matile, S. Rigid Oligonaphthalenediimide Rods as Transmembrane Anion−π Slides. J. Am. Chem. Soc. 2006, 128 (46), 14788−14789. (284) Li, J.; Pang, X.; Wang, Y.; Che, Y.; Zhao, J. A New Insight into Fluoride Anion in Electron Transfer Reactions. Catal. Today 2014, 224, 258−262. (285) Fang, X.; Guo, M.-D.; Weng, L.-J.; Chen, Y.; Lin, M.-J. Cooperative Effect of anion-π and Electrostatic Interactions in NIR Absorbing Phenolate Naphthalene Diimide Conjugates. Dyes Pigm. 2015, 113, 251−256. (286) Zuffo, M.; Doria, F.; Spalluto, V.; Ladame, S.; Freccero, M. Red/NIR G-Quadruplex Sensing, Harvesting Blue Light by a Coumarin−Naphthalene Diimide Dyad. Chem. - Eur. J. 2015, 21 (49), 17596−17600. (287) Ajayakumar, M. R.; Asthana, D.; Mukhopadhyay, P. CoreModified Naphthalenediimides Generate Persistent Radical Anion and Cation: New Panchromatic NIR Probes. Org. Lett. 2012, 14 (18), 4822−4825. (288) Ghule, N. V.; Bhosale, R. S.; Puyad, A. L.; Bhosale, S. V.; Bhosale, S. V. Naphthalenediimide Amphiphile Based Colorimetric Probe for Recognition of Cu2+ and Fe3+ Ions. Sens. Actuators, B 2016, 227, 17−23. (289) Lu, X.; Zhu, W.; Xie, Y.; Li, X.; Gao, Y.; Li, F.; Tian, H. NearIR Core-Substituted Naphthalenediimide Fluorescent Chemosensors for Zinc Ions: Ligand Effects on PET and ICT Channels. Chem. - Eur. J. 2010, 16 (28), 8355−8364. (290) Pal, S.; Hatai, J.; Samanta, M.; Shaurya, A.; Bandyopadhyay, S. A Highly Selective Chemodosimeter for Fast Detection and Intracellular Imaging of Hg2+ Ions Based on a DithiocarbamateIsothiocyanate Conversion in Aqueous Ethanol. Org. Biomol. Chem. 2014, 12 (7), 1072−1078. 11787

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

(291) Li, Q.; Peng, M.; Li, H.; Zhong, C.; Zhang, L.; Cheng, X.; Peng, X.; Wang, Q.; Qin, J.; Li, Z. A New “Turn-on” Naphthalenedimide-Based Chemosensor for Mercury Ions with High Selectivity: Successful Utilization of the Mechanism of Twisted Intramolecular Charge Transfer, Near-IR Fluorescence, and Cell Images. Org. Lett. 2012, 14 (8), 2094−2097. (292) Ajayakumar, M. R.; Mukhopadhyay, P. Naphthalene-bishydrazimide: Radical Anions and ICT as new Bimodal Probes for Differential Sensing of a Library of Amines. Chem. Commun. 2009, 3702−3704. (293) Rombouts, J. A.; Ravensbergen, J.; Frese, R. N.; Kennis, J. T. M.; Ehlers, A. W.; Slootweg, J. C.; Ruijter, E.; Lammertsma, K.; Orru, R. V. A. Synthesis and Photophysics of a Red-Light Absorbing Supramolecular Chromophore System. Chem. - Eur. J. 2014, 20 (33), 10285−10291. (294) Ono, T.; Hatanaka, S.; Hisaeda, Y. Turn-on Fluorogenic and Chromogenic Detection of Small Aromatic Hydrocarbon Vapors by a Porous Supramolecular Host. Chem. - Eur. J. 2016, 22, 10346. (295) Martínez-Quiroz, M.; Ochoa-Terán, A.; Pina-Luis, G.; Santacruz Ortega, H. Photoinduced Electron Transfer in N,Nbis(pyridylmethyl)naphthalenediimides: Study of Their Potential as pH Chemosensors. Supramol. Chem. 2016, 1−8. (296) Zhou, C.; Li, Y.; Zhao, Y.; Zhang, J.; Yang, W.; Li, Y. An Unusual Addition Reaction for Constructing a Novel pH-Controlled Fluorescence Switch. Org. Lett. 2011, 13 (2), 292−295. (297) Cox, R. P.; Higginbotham, H. F.; Graystone, B. A.; Sandanayake, S.; Langford, S. J.; Bell, T. D. M. A New Fluorescent H+ Sensor Based on Core-Substituted Naphthalene Diimide. Chem. Phys. Lett. 2012, 521, 59−63. (298) Higginbotham, H. F.; Cox, R. P.; Sandanayake, S.; Graystone, B. A.; Langford, S. J.; Bell, T. D. M. A Fluorescent ″2 in 1″ Proton Sensor and Polarity Probe Based on Core Substituted Naphthalene Diimide. Chem. Commun. 2013, 49 (44), 5061−5063. (299) Bhosale, S. V.; Adsul, M.; Shitre, G. V.; Bobe, S. R.; Bhosale, S. V.; Privér, S. H. A Pyridyl-Monoannulated Naphthalene Diimide Motif Self-Assembles into Tuneable Nanostructures by Means of Solvophobic Control. Chem. - Eur. J. 2013, 19 (23), 7310−7313. (300) Ghule, N. V.; Bhosale, R. S.; Kharat, K.; Puyad, A. L.; Bhosale, S. V.; Bhosale, S. V. A Naphthalenediimide-Based Fluorescent Sensor for Detecting the pH within the Rough Endoplasmic Reticulum of Living Cells. ChemPlusChem 2015, 80 (3), 485−489. (301) Doria, F.; Folini, M.; Grande, V.; Cimino-Reale, G.; Zaffaroni, N.; Freccero, M. Naphthalene Diimides as Red Fluorescent pH Sensors for Functional Cell Imaging. Org. Biomol. Chem. 2015, 13 (2), 570−576. (302) Doria, F.; Nadai, M.; Sattin, G.; Pasotti, L.; Richter, S. N.; Freccero, M. Water Soluble Extended Naphthalene Diimides as pH Fluorescent Sensors and G-Quadruplex Ligands. Org. Biomol. Chem. 2012, 10 (19), 3830−3840. (303) Doria, F.; Gallati, C. M.; Freccero, M. Hydrosoluble and Solvatochromic Naphthalene Diimides with NIR Absorption. Org. Biomol. Chem. 2013, 11 (45), 7838−7842. (304) Ghule, N. V.; Kharat, K.; Bhosale, R. S.; Puyad, A. L.; Bhosale, S. V.; Bhosale, S. V. The Fluorescence Detection of Autophagosomes in Live Cells under Starvation using Core-substituted Naphthalenediimide Probes. RSC Adv. 2016, 6 (2), 1644−1648. (305) Doria, F.; Amendola, V.; Grande, V.; Bergamaschi, G.; Freccero, M. Naphthalene Diimides as Selective Naked-eye Chemosensor for Copper(II) in Aqueous Solution. Sens. Actuators, B 2015, 212, 137−144. (306) Chen, D.; Avestro, A.-J.; Chen, Z.; Sun, J.; Wang, S.; Xiao, M.; Erno, Z.; Algaradah, M. M.; Nassar, M. S.; Amine, K.; Meng, Y.; Stoddart, J. F. A Rigid Naphthalenediimide Triangle for Organic Rechargeable Lithium-Ion Batteries. Adv. Mater. 2015, 27 (18), 2907− 2912. (307) Vadehra, G. S.; Maloney, R. P.; Garcia-Garibay, M. A.; Dunn, B. Naphthalene Diimide Based Materials with Adjustable Redox Potentials: Evaluation for Organic Lithium-Ion Batteries. Chem. Mater. 2014, 26 (24), 7151−7157.

(308) Mizuno, A.; Shuku, Y.; Suizu, R.; Matsushita, M. M.; Tsuchiizu, M.; Reta Mañeru, D.; Illas, F.; Robert, V.; Awaga, K. Discovery of the K4 Structure Formed by a Triangular π Radical Anion. J. Am. Chem. Soc. 2015, 137 (24), 7612−7615. (309) Liu, Z.; Sun, J.; Zhou, Y.; Zhang, Y.; Wu, Y.; Nalluri, S. K. M.; Wang, Y.; Samanta, A.; Mirkin, C. A.; Schatz, G. C.; Stoddart, J. F. Supramolecular Gelation of Rigid Triangular Macrocycles through Rings of Multiple C−H···O Interactions Acting Cooperatively. J. Org. Chem. 2016, 81 (6), 2581−2588. (310) Veerababu, M.; Varadaraju, U. V.; Kothandaraman, R. Reversible Lithium Storage Behaviour of Aromatic Diimide Dilithium Carboxylates. Electrochim. Acta 2016, 193, 80−87. (311) Zhou, D.; Wang, Y.; Jia, J.; Yu, W.; Qu, B.; Li, X.; Sun, X. HBonding and Charging Mediated Aggregation and Emission for Fluorescence Turn-on Detection of Hydrazine Hydrate. Chem. Commun. 2015, 51 (53), 10656−10659. (312) Zong, L.; Song, Y.; Li, Q.; Li, Z. A “turn-on” Fluorescence Probe Towards Copper Ions Based on Core-substitued Naphthalene Diimide. Sens. Actuators, B 2016, 226, 239−244. (313) Li, N.; Zong, L.; Li, Q.; Li, Z. An Imidazole-Containing CoreSubstituted Naphthalene Diimide: Fluorescent Sensing Properties Toward Copper Ion and Optimized Selectivity by Tuning the Solvent Medium. Sens. Actuators, B 2015, 207 (Part A), 827−832. (314) Lasitha, P.; Prasad, E. Orange Red Emitting Naphthalene Diimide Derivative Containing Dendritic Wedges: Aggregation Induced Emission (AIE) and Detection of Picric Acid (PA). RSC Adv. 2015, 5 (52), 41420−41427. (315) Li, G.-B.; Liao, Z.-H.; Pan, R.-K.; Liu, J.-M.; Liu, S.-G.; Zhang, Z. Synthesis, Fluorescence, and Sorption Properties of Cobalt Coordination Polymers of the N,N′-bis(4-pyridylmethyl)naphthalene diimide Ligand. Transition Met. Chem. 2015, 40 (6), 691−697. (316) Kübel, J.; Schroot, R.; Wächtler, M.; Schubert, U. S.; Dietzek, B.; Jäger, M. Photoredox-active Dyads Based on a Ru(II) Photosensitizer Equipped with Electron Donor or Acceptor Polymer Chains: A Spectroscopic Study of Light-Induced Processes toward Efficient Charge Separation. J. Phys. Chem. C 2015, 119 (9), 4742−4751. (317) Pandeeswar, M.; Govindaraju, T. Bioinspired Nanoarchitectonics of Naphthalene Diimide to Access 2D Sheets of Tunable Size, Shape, and Optoelectronic Properties. J. Inorg. Organomet. Polym. Mater. 2015, 25 (2), 293−300. (318) Black, S. P.; Wood, D. M.; Schwarz, F. B.; Ronson, T. K.; Holstein, J. J.; Stefankiewicz, A. R.; Schalley, C. A.; Sanders, J. K. M.; Nitschke, J. R. Catenation and Encapsulation Induce Distinct Reconstitutions within a Dynamic Library of Mixed-Ligand Zn4L6 Cages. Chem. Sci. 2016, 7, 2614. (319) Jazwinski, J.; Blacker, A. J.; Lehn, J.-M.; Cesario, M.; Guilhem, J.; Pascard, C. Cyclo-Bisintercalands: Synthesis and Structure of an Intercalative Inclusion Complex, and Anion Binding Properties. Tetrahedron Lett. 1987, 28 (48), 6060. (320) Zych, A. J.; Iverson, B. L. Synthesis and Conformational Characterization of Tethered, Self-Complexing 1,5-Dialkoxynaphthalene/1,4,5,8-Naphthalenetetracarboxylic Diimide Systems. J. Am. Chem. Soc. 2000, 122 (37), 8898−8909. (321) Johnstone, K. D.; Bampos, N.; Sanders, J. K. M.; Gunter, M. J. Gel-Phase HR-MAS 1H NMR Spectroscopy as a Probe for SolidTethered Diimide Rotaxanes and Catenanes. New J. Chem. 2006, 30 (6), 861−867. (322) Hagihara, S.; Gremaud, L.; Bollot, G.; Mareda, J.; Matile, S. Screening of π-Basic Naphthalene and Anthracene Amplifiers for πAcidic Synthetic Pore Sensors. J. Am. Chem. Soc. 2008, 130 (13), 4347−4351. (323) Fioravanti, G.; Haraszkiewicz, N.; Kay, E. R.; Mendoza, S. M.; Bruno, C.; Marcaccio, M.; Wiering, P. G.; Paolucci, F.; Rudolf, P.; Brouwer, A. M.; Leigh, D. A. Three State Redox-Active Molecular Shuttle That Switches in Solution and on a Surface. J. Am. Chem. Soc. 2008, 130 (8), 2593−2601. (324) Baggerman, J.; Haraszkiewicz, N.; Wiering, P. G.; Fioravanti, G.; Marcaccio, M.; Paolucci, F.; Kay, E. R.; Leigh, D. A.; Brouwer, A. 11788

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Aromatic Diimide Acceptor Units. Photochem. Photobiol. 2006, 82 (6), 1530−1540. (344) Bullock, J. E.; Vagnini, M. T.; Ramanan, C.; Co, D. T.; Wilson, T. M.; Dicke, J. W.; Marks, T. J.; Wasielewski, M. R. Photophysics and Redox Properties of Rylene Imide and Diimide Dyes Alkylated Ortho to the Imide Groups. J. Phys. Chem. B 2010, 114 (5), 1794−1802. (345) Adiga, S. P.; Shukla, D. Electronic Structure and ChargeTransport Properties of N,N′-Bis(cyclohexyl)naphthalene Diimide. J. Phys. Chem. C 2010, 114 (6), 2751−2755. (346) Tanaka, H.; Litvinchuk, S.; Tran, D.-H.; Bollot, G.; Mareda, J.; Sakai, N.; Matile, S. Adhesive π-Clamping within Synthetic Multifunctional Pores. J. Am. Chem. Soc. 2006, 128 (50), 16000−16001. (347) Steinberg-Yfrach, G.; Liddell, P. A.; Hung, S.-C.; Moore, A. L.; Gust, D.; Moore, T. A. Conversion of Light Energy to Proton Potential in Liposomes by Artificial Photosynthetic Reaction Centres. Nature 1997, 385 (6613), 239−241. (348) Osuka, A.; Zhang, R.-P.; Maruyama, K.; Ohno, T.; Nozaki, K. Synthesis and Intramolecular Charge Separation of Fixed-Distance Triads Consisting of Zinc Porphyrin, Metal-Free Porphyrin, and Electron-Accepting Diimide Moiety. Bull. Chem. Soc. Jpn. 1993, 66 (12), 3773−3782. (349) Shiratori, H.; Ohno, T.; Nozaki, K.; Yamazaki, I.; Nishimura, Y.; Osuka, A. Coordination Control of Intramolecular Electron Transfer in Boronate-Bridged Zinc PorphyrinDiimide Molecules. J. Org. Chem. 2000, 65 (25), 8747−8757. (350) Gosztola, D.; Niemczyk, M. P.; Wasielewski, M. R. Picosecond Molecular Switch Based on Bidirectional Inhibition of Photoinduced Electron Transfer Using Photogenerated Electric Fields. J. Am. Chem. Soc. 1998, 120 (20), 5118−5119. (351) Heinen, U.; Berthold, T.; Kothe, G.; Stavitski, E.; Galili, T.; Levanon, H.; Wiederrecht, G.; Wasielewski, M. R. High Time Resolution Q-Band EPR Study of Sequential Electron Transfer in a Triad Oriented in a Liquid Crystal. J. Phys. Chem. A 2002, 106 (10), 1933−1937. (352) Sessler, J. L.; Brown, C. T.; O’Connor, D.; Springs, S. L.; Wang, R.; Sathiosatham, M.; Hirose, T. A Rigid Chlorin−Naphthalene Diimide Conjugate. A Possible New Noncovalent Electron Transfer Model System. J. Org. Chem. 1998, 63 (21), 7370−7374. (353) Osuka, A.; Yoneshima, R.; Shiratori, H.; Osuka, A.; Okada, T.; Taniguchi, S.; Mataga, N. Electron Transfer in a Hydrogen-Bonded Assembly Consisting of Porphyrin-Diimide. Chem. Commun. 1998, 1567−1568. (354) Ghiggino, K. P.; Hutchison, J. A.; Langford, S. J.; Latter, M. J.; Lee, M. A.; Takezaki, M. A Simple Dyad Exhibiting Microsecond Charge-Separation in Non-Polar Solvents. Aust. J. Chem. 2006, 59 (3), 179−185. (355) Saito, K.; Kashiwagi, Y.; Ohkubo, K.; Fukuzumi, S. An Extremely Long-Lived Charge-Separated State of Zinc Tetraphenylporphyrin Coordinated with Pyridylnaphthalene-Diimide. J. Porphyrins Phthalocyanines 2006, 10 (12), 1371−1379. (356) Chaignon, F.; Falkenstrom, M.; Karlsson, S.; Blart, E.; Odobel, F.; Hammarstrom, L. Very Large Acceleration of the Photoinduced Electron Transfer in a Ru(bpy)3-Naphthalene Bisimide Dyad Bridged on the Naphthyl Core. Chem. Commun. 2007, 64−66. (357) Banerji, N.; Fürstenberg, A.; Bhosale, S.; Sisson, A. L.; Sakai, N.; Matile, S.; Vauthey, E. Ultrafast Photoinduced Charge Separation in Naphthalene Diimide Based Multichromophoric Systems in Liquid Solutions and in a Lipid Membrane. J. Phys. Chem. B 2008, 112 (30), 8912−8922. (358) Banerji, N.; Duvanel, G.; Perez-Velasco, A.; Maity, S.; Sakai, N.; Matile, S.; Vauthey, E. Excited-State Dynamics of Hybrid Multichromophoric Systems: Toward an Excitation Wavelength Control of the Charge Separation Pathways. J. Phys. Chem. A 2009, 113 (29), 8202−8212. (359) Bhosale, R.; Misek, J.; Sakai, N.; Matile, S. Supramolecular n/pHeterojunction Photosystems with Oriented Multicolored Antiparallel Redox Gradients (OMARG-SHJs). Chem. Soc. Rev. 2010, 39 (1), 138− 149.

M. Induction of Motion in a Synthetic Molecular Machine: Effect of Tuning the Driving Force. Chem. - Eur. J. 2013, 19 (18), 5566−5577. (325) Jacquot de Rouville, H.-P.; Iehl, J.; Bruns, C. J.; McGrier, P. L.; Frasconi, M.; Sarjeant, A. A.; Stoddart, J. F. A Neutral Naphthalene Diimide [2]Rotaxane. Org. Lett. 2012, 14 (20), 5188−5191. (326) Avestro, A.-J.; Gardner, D. M.; Vermeulen, N. A.; Wilson, E. A.; Schneebeli, S. T.; Whalley, A. C.; Belowich, M. E.; Carmieli, R.; Wasielewski, M. R.; Stoddart, J. F. Gated Electron Sharing Within Dynamic Naphthalene Diimide-Based Oligorotaxanes. Angew. Chem., Int. Ed. 2014, 53 (17), 4442−4449. (327) Kasai, Y.; Sakamoto, C.; Muroya, N.; Kato, S.-i.; Nakamura, Y. Synthesis of [60]Fullerene-Containing [2]Rotaxanes using Axle Molecules Bearing Donor Moiety. Tetrahedron Lett. 2011, 52 (5), 623−625. (328) Chen, L.; Zhang, Y.-M.; Wang, L.-H.; Liu, Y. Molecular Binding Behaviors of Pyromellitic and Naphthalene Diimide Derivatives by Tetrasulfonated 1,5-Dinaphtho-(3n+8)-crown-n (n = 8, 10) in Aqueous Solution. J. Org. Chem. 2013, 78 (11), 5357−5363. (329) Jiang, Q.; Zhang, H.-Y.; Han, M.; Ding, Z.-J.; Liu, Y. pHControlled Intramolecular Charge-Transfer Behavior in Bistable [3]Rotaxane. Org. Lett. 2010, 12 (8), 1728−1731. (330) Müllen, K. M.; Davis, J. J.; Beer, P. D. Anion Induced Displacement Studies in Naphthalene Diimide Containing Interpenetrated and Interlocked Structures. New J. Chem. 2009, 33 (4), 769−776. (331) Mandal, A. K.; Suresh, M.; Das, A. Studies on [3]Pseudorotaxane Formation from a Bis-azacrown Derivative as Host and Imidazolium Ion-derivatives as Guest. Org. Biomol. Chem. 2011, 9 (13), 4811−4817. (332) Yang, W.; Li, Y.; Zhang, J.; Chen, N.; Chen, S.; Liu, H.; Li, Y. Directed Synthesis of [2]Catenanes Incorporating Naphthalenediimide and Crown Ethers by Associated Interactions of Templates. J. Org. Chem. 2011, 76 (19), 7750−7756. (333) Ponnuswamy, N.; Cougnon, F. B. L.; Pantoş, G. D.; Sanders, J. K. M. Homochiral and meso Figure Eight Knots and a Solomon Link. J. Am. Chem. Soc. 2014, 136 (23), 8243−8251. (334) Jiang, W.; Han, M.; Zhang, H.-Y.; Zhang, Z.-J.; Liu, Y. A Double Plug−Socket System Capable of Molecular Keypad Locks through Controllable Photooxidation. Chem. - Eur. J. 2009, 15 (38), 9938−9945. (335) Tanaka, H.; Bollot, G.; Mareda, J.; Litvinchuk, S.; Tran, D.-H.; Sakai, N.; Matile, S. Synthetic Pores with Sticky π-Clamps. Org. Biomol. Chem. 2007, 5 (9), 1369−1380. (336) Gorteau, V.; Bollot, G.; Mareda, J.; Matile, S. Rigid-rod Anionπ Slides for Multiion Hopping Across Lipid Bilayers. Org. Biomol. Chem. 2007, 5 (18), 3000−3012. (337) Gorteau, V.; Julliard, M. D.; Matile, S. Hydrophilic Anchors for Transmembrane Anion−π Slides. J. Membr. Sci. 2008, 321 (1), 37−42. (338) Sakai, N.; Charbonnaz, P.; Ward, S.; Matile, S. Ion-Gated Synthetic Photosystems. J. Am. Chem. Soc. 2014, 136 (15), 5575− 5578. (339) Doval, D. A.; Fin, A.; Takahashi-Umebayashi, M.; Riezman, H.; Roux, A.; Sakai, N.; Matile, S. Amphiphilic Dynamic NDI and PDI Probes: Imaging Microdomains in Giant Unilamellar Vesicles. Org. Biomol. Chem. 2012, 10 (30), 6087−6093. (340) Widanapathirana, L.; Zhao, Y. Aromatically Functionalized Cyclic Tricholate Macrocycles: Aggregation, Transmembrane Pore Formation, Flexibility, and Cooperativity. J. Org. Chem. 2012, 77 (10), 4679−4687. (341) Wasielewski, M. R. Photoinduced Electron Transfer in Supramolecular Systems for Artificial Photosynthesis. Chem. Rev. 1992, 92 (3), 435−461. (342) Deisenhofer, J.; Epp, O.; Miki, K.; Huber, R.; Michel, H. Structure of the Protein Subunits in the Photosynthetic Reaction Centre of Rhodopseudomonas Viridis at 3Å Resolution. Nature 1985, 318 (6047), 618−624. (343) Langford, S. J.; Latter, M. J.; Woodward, C. P. Progress in Charge Transfer Systems Utilizing Porphyrin Donors and Simple 11789

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

S1 States as a Basis for a Molecular Switch. J. Phys. Chem. A 2010, 114 (4), 1709−1721. (378) Favereau, L.; Makhal, A.; Pellegrin, Y.; Blart, E.; Petersson, J.; Göransson, E.; Hammarström, L.; Odobel, F. A Molecular Tetrad That Generates a High-Energy Charge-Separated State by Mimicking the Photosynthetic Z-Scheme. J. Am. Chem. Soc. 2016, 138 (11), 3752−3760. (379) El-Khouly, M. E.; Kim, J. H.; Kay, K.-Y.; Choi, C. S.; Ito, O.; Fukuzumi, S. Synthesis and Photoinduced Intramolecular Processes of Light-Harvesting Silicon Phthalocyanine−Naphthalenediimide−Fullerene Connected Systems. Chem. - Eur. J. 2009, 15 (21), 5301−5310. (380) El-Khouly, M. E.; Wijesinghe, C. A.; Nesterov, V. N.; Zandler, M. E.; Fukuzumi, S.; D’Souza, F. Ultrafast Photoinduced Energy and Electron Transfer in Multi-Modular Donor−Acceptor Conjugates. Chem. - Eur. J. 2012, 18 (43), 13844−13853. (381) Guo, S.; Sun, J.; Ma, L.; You, W.; Yang, P.; Zhao, J. Visible Light-Harvesting Naphthalenediimide (NDI)-C60 Dyads as HeavyAtom-Free Organic Triplet Photosensitizers for Triplet−Triplet Annihilation Based Upconversion. Dyes Pigm. 2013, 96 (2), 449−458. (382) Poddutoori, P. K.; Zarrabi, N.; Moiseev, A. G.; Gumbau-Brisa, R.; Vassiliev, S.; van der Est, A. Long-Lived Charge Separation in Novel Axial Donor−Porphyrin−Acceptor Triads Based on Tetrathiafulvalene, Aluminum(III) Porphyrin and Naphthalenediimide. Chem. Eur. J. 2013, 19 (9), 3148−3161. (383) Supur, M.; El-Khouly, M. E.; Seok, J. H.; Kay, K.-Y.; Fukuzumi, S. Elongation of Lifetime of the Charge-Separated State of Ferrocene− Naphthalenediimide−[60]Fullerene Triad via Stepwise Electron Transfer. J. Phys. Chem. A 2011, 115 (50), 14430−14437. (384) El-Khouly, M. E.; Kim, J.-H.; Kim, J.-H.; Kay, K.-Y.; Fukuzumi, S. Subphthalocyanines as Light-Harvesting Electron Donor and Electron Acceptor in Artificial Photosynthetic Systems. J. Phys. Chem. C 2012, 116 (37), 19709−19717. (385) Lambert, C.; Wagener, R.; Klein, J. H.; Grelaud, G.; Moos, M.; Schmiedel, A.; Holzapfel, M.; Bruhn, T. A Photoinduced MixedValence State in an Organic bis-Triarylamine Mixed-Valence Compound with an Iridium-Metal-Bridge. Chem. Commun. 2014, 50 (77), 11350−11353. (386) Klein, J. H.; Schmidt, D.; Steiner, U. E.; Lambert, C. Complete Monitoring of Coherent and Incoherent Spin Flip Domains in the Recombination of Charge-Separated States of Donor-Iridium Complex-Acceptor Triads. J. Am. Chem. Soc. 2015, 137 (34), 11011−11021. (387) Kaiser, C.; Schmiedel, A.; Holzapfel, M.; Lambert, C. LongLived Singlet and Triplet Charge Separated States in Small Cyclophane-Bridged Triarylamine−Naphthalene Diimide Dyads. J. Phys. Chem. C 2012, 116 (29), 15265−15280. (388) Zieschang, F.; Schreck, M. H.; Schmiedel, A.; Holzapfel, M.; Klein, J. H.; Walter, C.; Engels, B.; Lambert, C. Photoinduced Electron Transfer Dynamics in Triarylamine−Naphthalene Diimide Cascades. J. Phys. Chem. C 2014, 118 (48), 27698−27714. (389) Röger, C.; Miloslavina, Y.; Brunner, D.; Holzwarth, A. R.; Würthner, F. Self-Assembled Zinc Chlorin Rod Antennae Powered by Peripheral Light-Harvesting Chromophores. J. Am. Chem. Soc. 2008, 130 (18), 5929−5939. (390) Tu, S.; Kim, S. H.; Joseph, J.; Modarelli, D. A.; Parquette, J. R. Self-Assembly of a Donor−Acceptor Nanotube. A Strategy To Create Bicontinuous Arrays. J. Am. Chem. Soc. 2011, 133 (47), 19125−19130. (391) Tu, S.; Kim, S. H.; Joseph, J.; Modarelli, D. A.; Parquette, J. R. Proton-Coupled Self-Assembly of a Porphyrin-Naphthalenediimide Dyad. ChemPhysChem 2013, 14 (8), 1609−1617. (392) Mondal, T.; Sakurai, T.; Yoneda, S.; Seki, S.; Ghosh, S. Semiconducting Nanotubes by Intrachain Folding Following Macroscopic Assembly of a Naphthalene−Diimide (NDI) Appended Polyurethane. Macromolecules 2015, 48 (4), 879−888. (393) Gujrati, M. D.; Kumar, N. S. S.; Brown, A. S.; Captain, B.; Wilson, J. N. Luminescent Charge-Transfer Complexes: Tuning Emission in Binary Fluorophore Mixtures. Langmuir 2011, 27 (11), 6554−6558. (394) Tan, L.; Zhang, G.; Zhang, D.; Zhu, D. Linear and Cyclic Tetrathiafulvalene−Naphthalenediimide Donor−Acceptor Molecules:

(360) Porus, M.; Maroni, P.; Bhosale, R.; Sakai, N.; Matile, S.; Borkovec, M. Zipper and Layer-by-Layer Assemblies of Artificial Photosystems Analyzed by Combining Optical and Piezoelectric Surface Techniques. Langmuir 2011, 27 (11), 7213−7221. (361) Miros, F. N.; Matile, S. Core-Substituted Naphthalenediimides: LUMO Levels Revisited, in Comparison with Preylenediimides with Sulfur Redox Switches in the Core. ChemistryOpen 2016, 5, 219. (362) Mulder, J. R.; Guerra, C. F.; Slootweg, J. C.; Lammertsma, K.; Bickelhaupt, F. M. Substituent Effects on the Optical Properties of Naphthalenediimides: A Frontier Orbital Analysis Across the Periodic Table. J. Comput. Chem. 2016, 37 (2), 304−313. (363) Bhosale, R.; Perez-Velasco, A.; Ravikumar, V.; Kishore, R. S. K.; Kel, O.; Gomez-Casado, A.; Jonkheijm, P.; Huskens, J.; Maroni, P.; Borkovec, M.; Sawada, T.; Vauthey, E.; Sakai, N.; Matile, S. Topologically Matching Supramolecular n/p-Heterojunction Architectures. Angew. Chem., Int. Ed. 2009, 48 (35), 6461−6464. (364) Sao, S.; Samanta, B. R.; Chaudhuri, D. An Unusual OneDonor-Two-Acceptor Interaction in a Pair of Covalently Bridged Naphthalenediimide Dimers. RSC Adv. 2016, 6 (41), 34350−34353. (365) Lu, C.; Fujitsuka, M.; Sugimoto, A.; Majima, T. Unprecedented Intramolecular Electron Transfer from Excited Perylenediimide Radical Anion. J. Phys. Chem. C 2016, 120, 12734. (366) Jaggi, M.; Schmid, B.; Liu, S.-X.; Bhosale, S. V.; Rivadehi, S.; Langford, S. J.; Decurtins, S. A Tetrathiafulvalene-Functionalized Naphthalene Diimide: Synthesis, Electrochemical and Photophysical Properties. Tetrahedron 2011, 67 (38), 7231−7235. (367) Banerji, N.; Bhosale, S. V.; Petkova, I.; Langford, S. J.; Vauthey, E. Ultrafast Excited-State Dynamics of Strongly Coupled Porphyrin/ Core-substituted-Naphthalenediimide Dyads. Phys. Chem. Chem. Phys. 2011, 13 (3), 1019−1029. (368) Handayani, M.; Gohda, S.; Tanaka, D.; Ogawa, T. Design and Synthesis of Perpendicularly Connected Metal Porphyrin−Imide Dyads for Two-Terminal Wired Single Molecular Diodes. Chem. Eur. J. 2014, 20 (25), 7655−7664. (369) Villamaina, D.; Bhosale, S. V.; Langford, S. J.; Vauthey, E. Excited-State Dynamics of Porphyrin-Naphthalenediimide-Porphyrin Triads. Phys. Chem. Chem. Phys. 2013, 15 (4), 1177−1187. (370) Tong, L. H.; Pengo, P.; Clegg, W.; Lowe, J. P.; Raithby, P. R.; Sanders, J. K. M.; Pascu, S. I. Complexes of Aryl-Substituted Porphyrins and Naphthalenediimide (NDI): Investigations by Synchrotron X-ray Diffraction and NMR Spectroscopy. Dalton Trans. 2011, 40 (41), 10833−10842. (371) Villamaina, D.; Kelson, M. M. A.; Bhosale, S. V.; Vauthey, E. Excitation Wavelength Dependence of the Charge Separation Pathways in Tetraporphyrin-Naphthalene Diimide Pentads. Phys. Chem. Chem. Phys. 2014, 16 (11), 5188−5200. (372) Yushchenko, O.; Hangarge, R. V.; Mosquera-Vazquez, S.; Boshale, S. V.; Vauthey, E. Electron, Hole, Singlet, and Triplet Energy Transfer in Photoexcited Porphyrin-Naphthalenediimide Dyads. J. Phys. Chem. B 2015, 119 (24), 7308−7320. (373) Robotham, B.; Lastman, K. A.; Langford, S. J.; Ghiggino, K. P. Ultrafast Electron Transfer in a Porphyrin-Amino Naphthalene Diimide Dyad. J. Photochem. Photobiol., A 2013, 251, 167−174. (374) Roznyatovskiy, V. V.; Gardner, D. M.; Eaton, S. W.; Wasielewski, M. R. Radical Anions of Trifluoromethylated Perylene and Naphthalene Imide and Diimide Electron Acceptors. Org. Lett. 2014, 16 (3), 696−699. (375) Maniam, S.; Higginbotham, H. F.; Guo, S.-X.; Bell, T. D. M.; Izgorodina, E. I.; Langford, S. J. A Redox Switchable Dihydrobenzo[b]pyrazine Push-Pull System. Asian J. Org. Chem. 2014, 3 (5), 619− 623. (376) Young, E. R.; Rosenthal, J.; Hodgkiss, J. M.; Nocera, D. G. Comparative PCET Study of a Donor−Acceptor Pair Linked by Ionized and Nonionized Asymmetric Hydrogen-Bonded Interfaces. J. Am. Chem. Soc. 2009, 131 (22), 7678−7684. (377) Wallin, S.; Monnereau, C.; Blart, E.; Gankou, J.-R.; Odobel, F.; Hammarström, L. State-Selective Electron Transfer in an Unsymmetric Acceptor−Zn(II)porphyrin−Acceptor Triad: Toward a Controlled Directionality of Electron Transfer from the Porphyrin S2 and 11790

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Metal Ions-Promoted Electron Transfer. J. Org. Chem. 2011, 76 (21), 9046−9052. (395) Rananaware, A.; Samanta, M.; Bhosale, R. S.; Kobaisi, M. A.; Roy, B.; Bheemireddy, V.; Bhosale, S. V.; Bandyopadhyay, S.; Bhosale, S. V. Photomodulation of Fluoride Ion Binding Through Anion-π Interactions using a Photoswitchable Azobenzene System. Sci. Rep. 2016, 6, 22928. (396) Ajayakumar, M. R.; Hundal, G.; Mukhopadhyay, P. A Tetrastable Naphthalenediimide: Anion Induced Charge Transfer, Single and Double Electron Transfer for Combinational Logic Gates. Chem. Commun. 2013, 49 (70), 7684−7686. (397) Asthana, D.; Ajayakumar, M. R.; Pant, R. P.; Mukhopadhyay, P. NTCDA-TTF First Axial Fusion: Emergent Panchromatic, NIR Optical, Multi-State Redox and high Optical Contrast photooxidation. Chem. Commun. 2012, 48 (52), 6475−6477. (398) Liu, Y.; Wu, W.; Zhao, J.; Zhang, X.; Guo, H. Accessing the Long-Lived Near-IR-Emissive Triplet Excited State in Naphthalenediimide with Light-Harvesting Diimine Platinum(II) Bisacetylide Complex and its Application for Upconversion. Dalton Trans. 2011, 40 (36), 9085−9089. (399) Ma, L.; Guo, S.; Sun, J.; Zhang, C.; Zhao, J.; Guo, H. Green Light-Excitable Naphthalenediimide Acetylide-Containing Cyclometalated Ir(III) Complex with Long-Lived Triplet Excited States as Triplet Photosensitizers for Triplet-Triplet Annihilation Upconversion. Dalton Trans. 2013, 42 (18), 6478−6488. (400) Liu, L.; Guo, S.; Ma, J.; Xu, K.; Zhao, J.; Zhang, T. Broadband Visible-Light-Harvesting trans-Bis(alkylphosphine) Platinum(II)-Alkynyl Complexes with Singlet Energy Transfer between BODIPY and Naphthalene Diimide Ligands. Chem. - Eur. J. 2014, 20 (44), 14282−14295. (401) Guo, S.; Wu, W.; Guo, H.; Zhao, J. Room-Temperature LongLived Triplet Excited States of Naphthalenediimides and Their Applications as Organic Triplet Photosensitizers for Photooxidation and Triplet−Triplet Annihilation Upconversions. J. Org. Chem. 2012, 77 (8), 3933−3943. (402) Doria, F.; Manet, I.; Grande, V.; Monti, S.; Freccero, M. WaterSoluble Naphthalene Diimides as Singlet Oxygen Sensitizers. J. Org. Chem. 2013, 78 (16), 8065−8073. (403) Song, Q.; Li, F.; Wang, Z.; Zhang, X. A Supramolecular Strategy for Tuning the Energy Level of Naphthalenediimide: Promoted Formation of Radical Anions with Extraordinary Stability. Chem. Sci. 2015, 6 (6), 3342−3346. (404) Kelson, M. M. A.; Bhosale, R. S.; Ohkubo, K.; Jones, L. A.; Bhosale, S. V.; Gupta, A.; Fukuzumi, S.; Bhosale, S. V. A Simple ZincPorphyrin-NDI Dyad System Generates a Light Energy to Proton Potential Across a Lipid Membrane. Dyes Pigm. 2015, 120, 340−346. (405) Polizzi, N. F.; Eibling, M. J.; Perez-Aguilar, J. M.; Rawson, J.; Lanci, C. J.; Fry, H. C.; Beratan, D. N.; Saven, J. G.; Therien, M. J. Photoinduced Electron Transfer Elicits a Change in the Static Dielectric Constant of a de Novo Designed Protein. J. Am. Chem. Soc. 2016, 138 (7), 2130−2133. (406) Liu, J.-J.; Wang, Y.; Lin, M.-J.; Huang, C.-C.; Dai, W.-X. Photogeneration of two Reduction-Active Charge-Separated States in a Hybrid Crystal of Polyoxometalates and Naphthalene Diimides. Dalton Trans. 2015, 44 (2), 484−487. (407) Urbach, A. R. DNA Complexes: Durable Binders. Nat. Chem. 2011, 3 (11), 836−837. (408) Collie, G. W.; Promontorio, R.; Hampel, S. M.; Micco, M.; Neidle, S.; Parkinson, G. N. Structural Basis for Telomeric GQuadruplex Targeting by Naphthalene Diimide Ligands. J. Am. Chem. Soc. 2012, 134 (5), 2723−2731. (409) Marchetti, C.; Minarini, A.; Tumiatti, V.; Moraca, F.; Parrotta, L.; Alcaro, S.; Rigo, R.; Sissi, C.; Gunaratnam, M.; Ohnmacht, S. A.; Neidle, S.; Milelli, A. Macrocyclic Naphthalene Diimides as GQuadruplex Binders. Bioorg. Med. Chem. 2015, 23 (13), 3819−3830. (410) Sato, S.; Tsueda, M.; Takenaka, S. Electrochemical Detection of Aberrant Methylated Gene using Naphthalene Diimide Derivative Carrying four Ferrocene Moieties. J. Organomet. Chem. 2010, 695 (15−16), 1858−1862.

(411) Himuro, T.; Sato, S.; Takenaka, S.; Yasuda, T. Formation and Electrical Evaluation of a Single Metallized DNA Nanowire in a Nanochannel. Electroanalysis 2016, 28, 1448. (412) Doria, F.; Nadai, M.; Folini, M.; Di Antonio, M.; Germani, L.; Percivalle, C.; Sissi, C.; Zaffaroni, N.; Alcaro, S.; Artese, A.; Richter, S. N.; Freccero, M. Hybrid Ligand-Alkylating Agents Targeting Telomeric G-Quadruplex Structures. Org. Biomol. Chem. 2012, 10 (14), 2798−2806. (413) Nadai, M.; Cimino-Reale, G.; Sattin, G.; Doria, F.; Butovskaya, E.; Zaffaroni, N.; Freccero, M.; Palumbo, M.; Richter, S. N.; Folini, M. Assessment of Gene Promoter G-Quadruplex Binding and Modulation by a Naphthalene Diimide Derivative in Tumor Cells. Int. J. Oncol. 2014, 46 (1), 369−380. (414) Di Antonio, M.; Doria, F.; Richter, S. N.; Bertipaglia, C.; Mella, M.; Sissi, C.; Palumbo, M.; Freccero, M. Quinone Methides Tethered to Naphthalene Diimides as Selective G-Quadruplex Alkylating Agents. J. Am. Chem. Soc. 2009, 131 (36), 13132−13141. (415) Tumiatti, V.; Milelli, A.; Minarini, A.; Micco, M.; Gasperi Campani, A.; Roncuzzi, L.; Baiocchi, D.; Marinello, J.; Capranico, G.; Zini, M.; Stefanelli, C.; Melchiorre, C. Design, Synthesis, and Biological Evaluation of Substituted Naphthalene Imides and Diimides as Anticancer Agent∞. J. Med. Chem. 2009, 52 (23), 7873−7877. (416) Wang, Y.; Zhang, X.; Zhao, J.; Xie, S.; Wang, C. Nonhematotoxic Naphthalene Diimide Modified by Polyamine: Synthesis and Biological Evaluation. J. Med. Chem. 2012, 55 (7), 3502−3512. (417) Gunaratnam, M.; Swank, S.; Haider, S. M.; Galesa, K.; Reszka, A. P.; Beltran, M.; Cuenca, F.; Fletcher, J. A.; Neidle, S. Targeting Human Gastrointestinal Stromal Tumor Cells with a QuadruplexBinding Small Molecule. J. Med. Chem. 2009, 52 (12), 3774−3783. (418) Ohtsuka, K.; Komizo, K.; Takenaka, S. Synthesis and DNA Binding Behavior of a Naphthalene Diimide Derivative Carrying two Dicobalt Hexacarbonyl Complexes as an Infrared DNA Probe. J. Organomet. Chem. 2010, 695 (9), 1281−1286. (419) Hampel, S. M.; Sidibe, A.; Gunaratnam, M.; Riou, J.-F.; Neidle, S. Tetrasubstituted Naphthalene Diimide Ligands with Selectivity for Telomeric G-Quadruplexes and Cancer Cells. Bioorg. Med. Chem. Lett. 2010, 20 (22), 6459−6463. (420) Spinello, A.; Barone, G.; Grunenberg, J. Molecular Recognition of Naphthalene Diimide Ligands by Telomeric Quadruplex-DNA: the Importance of the Protonation State and Mediated Hydrogen Bonds. Phys. Chem. Chem. Phys. 2016, 18 (4), 2871−2877. (421) Gunaratnam, M.; Fuente, M. d. l.; Hampel, S. M.; Todd, A. K.; Reszka, A. P.; Schätzlein, A.; Neidle, S. Targeting Pancreatic Cancer with a G-Quadruplex Ligand. Bioorg. Med. Chem. 2011, 19 (23), 7151−7157. (422) McKnight, R. E.; Reisenauer, E.; Pintado, M. V.; Polasani, S. R.; Dixon, D. W. Substituent Effect on the Preferred DNA Binding Mode and Affinity of a Homologous Series of Naphthalene Diimides. Bioorg. Med. Chem. Lett. 2011, 21 (14), 4288−4291. (423) Xue, L.; Xi, H.; Kumar, S.; Gray, D.; Davis, E.; Hamilton, P.; Skriba, M.; Arya, D. P. Probing the Recognition Surface of a DNA Triplex: Binding Studies with Intercalator−Neomycin Conjugates. Biochemistry 2010, 49 (26), 5540−5552. (424) Rhoden Smith, A.; Ikkanda, B. A.; Holman, G. G.; Iverson, B. L. Subtle Recognition of 14-Base Pair DNA Sequences via Threading Polyintercalation. Biochemistry 2012, 51 (22), 4445−4452. (425) Holman, G. G.; Zewail-Foote, M.; Smith, A. R.; Johnson, K. A.; Iverson, B. L. A Sequence-Specific Threading Tetraintercalator with an Extremely Slow Dissociation Rate Constant. Nat. Chem. 2011, 3 (11), 875−881. (426) Ng, P.-S.; Laing, B. M.; Balasundarum, G.; Pingle, M.; Friedman, A.; Bergstrom, D. E. Synthesis and Evaluation of New Spacers for Use as dsDNA End-Caps. Bioconjugate Chem. 2010, 21 (8), 1545−1553. (427) Watanabe, S.; Ohtsuka, K.; Sato, S.; Takenaka, S. Discrimination of Phosphorylated Double Stranded DNA by Naphthalene Diimide having Zinc(II) Dipicolylamine Complexes. Bioorg. Med. Chem. 2011, 19 (3), 1361−1365. 11791

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

(428) Czerwinska, I.; Sato, S.; Takenaka, S. Improving the Affinity of Naphthalene Diimide Ligand to Telomeric DNA by Incorporating Zn2+ Ions into its Dipicolylamine Groups. Bioorg. Med. Chem. 2012, 20 (21), 6416−6422. (429) Sato, S.; Takenaka, S. Linker Effect of Ferrocenylnaphthalene Diimide Ligands in the Interaction with Double Stranded DNA. J. Organomet. Chem. 2008, 693 (7), 1177−1185. (430) Minarini, A.; Milelli, A.; Tumiatti, V.; Ferruzzi, L.; Marton, M.R.; Turrini, E.; Hrelia, P.; Fimognari, C. Design, Synthesis and Biological Evaluation of New Naphtalene Diimides Bearing Isothiocyanate Functionality. Eur. J. Med. Chem. 2012, 48, 124−131. (431) Milelli, A.; Tumiatti, V.; Micco, M.; Rosini, M.; Zuccari, G.; Raffaghello, L.; Bianchi, G.; Pistoia, V.; Fernando Díaz, J.; Pera, B.; Trigili, C.; Barasoain, I.; Musetti, C.; Toniolo, M.; Sissi, C.; Alcaro, S.; Moraca, F.; Zini, M.; Stefanelli, C.; Minarini, A. Structure−Activity Relationships of Novel Substituted Naphthalene Diimides as Anticancer Agents. Eur. J. Med. Chem. 2012, 57, 417−428. (432) Micco, M.; Collie, G. W.; Dale, A. G.; Ohnmacht, S. A.; Pazitna, I.; Gunaratnam, M.; Reszka, A. P.; Neidle, S. Structure-Based Design and Evaluation of Naphthalene Diimide G-Quadruplex Ligands As Telomere Targeting Agents in Pancreatic Cancer Cells. J. Med. Chem. 2013, 56 (7), 2959−2974. (433) Mpima, S.; Ohnmacht, S. A.; Barletta, M.; Husby, J.; Pett, L. C.; Gunaratnam, M.; Hilton, S. T.; Neidle, S. The Influence of Positional Isomerism on G-Quadruplex Binding and Anti-proliferative Activity of Tetra-substituted Naphthalene Diimide Compounds. Bioorg. Med. Chem. 2013, 21 (20), 6162−6170. (434) Sato, S.; Yamamura, K.; Takenaka, S. Naphthalene Diimide Carrying Two Cysteine Termini at Both Imide Linkers as a Molecular Staple. Electroanalysis 2013, 25 (8), 1831−1839. (435) Ikkanda, B. A.; Samuel, S. A.; Iverson, B. L. NDI and DAN DNA: Nucleic Acid-Directed Assembly of NDI and DAN. J. Org. Chem. 2014, 79 (5), 2029−2037. (436) Esaki, Y.; Islam, M. M.; Fujii, S.; Sato, S.; Takenaka, S. Design of Tetraplex Specific Ligands: Cyclic Naphthalene Diimide. Chem. Commun. 2014, 50 (45), 5967−5969. (437) Islam, M. M.; Fujii, S.; Sato, S.; Okauchi, T.; Takenaka, S. Thermodynamics and Kinetic Studies in the Binding Interaction of Cyclic Naphthalene Diimide Derivatives with Double stranded DNAs. Bioorg. Med. Chem. 2015, 23 (15), 4769−4776. (438) Komizo, K.; Ikedo, H.; Sato, S.; Takenaka, S. Metallization of Double-Stranded DNA Triggered by Bound Galactose-Modified Naphthalene Diimide. Bioconjugate Chem. 2014, 25 (8), 1547−1555. (439) Mokhir, A. A.; Kraemer, R. Conjugates of PNA with Naphthalene Diimide Derivatives Having a Broad Range of DNA Affinities. Bioconjugate Chem. 2003, 14 (5), 877−883. (440) Nadai, M.; Doria, F.; Germani, L.; Richter, S. N.; Freccero, M. A Photoreactive G-Quadruplex Ligand Triggered by Green Light. Chem. - Eur. J. 2015, 21 (6), 2330−2334. (441) Doria, F.; Oppi, A.; Manoli, F.; Botti, S.; Kandoth, N.; Grande, V.; Manet, I.; Freccero, M. A Naphthalene Diimide Dyad for Fluorescence Switch-on Detection of G-Quadruplexes. Chem. Commun. 2015, 51 (44), 9105−9108. (442) Hu, Z.; Arrowsmith, R. L.; Tyson, J. A.; Mirabello, V.; Ge, H.; Eggleston, I. M.; Botchway, S. W.; Dan Pantos, G.; Pascu, S. I. A Fluorescent Arg-Gly-Asp (RGD) Peptide-Naphthalenediimide (NDI) Conjugate for Imaging Integrin αvβ3 in vitro. Chem. Commun. 2015, 51 (32), 6901−6904. (443) Haubner, R.; Wester, H.-J.; Weber, W. A.; Mang, C.; Ziegler, S. I.; Goodman, S. L.; Senekowitsch-Schmidtke, R.; Kessler, H.; Schwaiger, M. Noninvasive Imaging of αvβ3 Integrin Expression Using 18F-labeled RGD-containing Glycopeptide and Positron Emission Tomography. Cancer Res. 2001, 61 (5), 1781−1785. (444) Prato, G.; Silvent, S.; Saka, S.; Lamberto, M.; Kosenkov, D. Thermodynamics of Binding of Di- and Tetrasubstituted Naphthalene Diimide Ligands to DNA G-Quadruplex. J. Phys. Chem. B 2015, 119 (8), 3335−3347. (445) Perrone, R.; Doria, F.; Butovskaya, E.; Frasson, I.; Botti, S.; Scalabrin, M.; Lago, S.; Grande, V.; Nadai, M.; Freccero, M.; Richter,

S. N. Synthesis, Binding and Antiviral Properties of Potent CoreExtended Naphthalene Diimides Targeting the HIV-1 Long Terminal Repeat Promoter G-Quadruplexes. J. Med. Chem. 2015, 58 (24), 9639−9652. (446) Gao, X.; Hu, Y. Development of n-type organic semiconductors for thin-film transistors: a viewpoint of molecular design. J. Mater. Chem. C 2014, 2 (17), 3099−3117. (447) Jiang, W.; Li, Y.; Wang, Z. Tailor-Made Rylene Arrays for High Performance n-Channel Semiconductors. Acc. Chem. Res. 2014, 47 (10), 3135−3147. (448) Mao, Z.; Le, T. P.; Vakhshouri, K.; Fernando, R.; Ruan, F.; Muller, E.; Gomez, E. D.; Sauvé, G. Processing Additive Suppresses Phase Separation in the Active Layer of Organic Photovoltaics based on Naphthalene Diimide. Org. Electron. 2014, 15 (11), 3384−3391. (449) Russ, B.; Robb, M. J.; Popere, B. C.; Perry, E. E.; Mai, C.-K.; Fronk, S. L.; Patel, S. N.; Mates, T. E.; Bazan, G. C.; Urban, J. J.; Chabinyc, M. L.; Hawker, C. J.; Segalman, R. A. Tethered Tertiary Amines as Solid-State n-Type Dopants for Solution-Processable Organic Semiconductors. Chem. Sci. 2016, 7 (3), 1914−1919. (450) Fujitsuka, M.; Kim, S. S.; Lu, C.; Tojo, S.; Majima, T. Intermolecular and Intramolecular Electron Transfer Processes from Excited Naphthalene Diimide Radical Anions. J. Phys. Chem. B 2015, 119 (24), 7275−7282. (451) Kalita, A.; Subbarao, N. V. V; Iyer, P. K. Large-Scale Molecular Packing and Morphology-Dependent High Performance Organic Field-Effect Transistor by Symmetrical Naphthalene Diimide Appended with Methyl Cyclohexane. J. Phys. Chem. C 2015, 119 (22), 12772−12779. (452) Wang, W.; Yuan, J.; Shi, G.; Zhu, X.; Shi, S.; Liu, Z.; Han, L.; Wang, H.-Q.; Ma, W. Inverted Planar Heterojunction Perovskite Solar Cells Employing Polymer as the Electron Conductor. ACS Appl. Mater. Interfaces 2015, 7 (7), 3994−3999. (453) Giussani, E.; Brambilla, L.; Fazzi, D.; Sommer, M.; Kayunkid, N.; Brinkmann, M.; Castiglioni, C. Structural Characterization of Highly Oriented Naphthalene-Diimide-Bithiophene Copolymer Films via Vibrational Spectroscopy. J. Phys. Chem. B 2015, 119 (5), 2062− 2073. (454) Alvey, P. M.; Reczek, J. J.; Lynch, V.; Iverson, B. L. A Systematic Study of Thermochromic Aromatic Donor−Acceptor Materials. J. Org. Chem. 2010, 75 (22), 7682−7690. (455) Suraru, S. L.; Würthner, F. Core-Tetrasubstituted Naphthalene Diimides by Stille Cross-Coupling Reactions and Characterization of their Optical and Redox Properties. Synthesis 2009, 2009 (11), 1841− 1845. (456) Shukla, D.; Nelson, S. F.; Freeman, D. C.; Rajeswaran, M.; Ahearn, W. G.; Meyer, D. M.; Carey, J. T. Thin-Film Morphology Control in Naphthalene-Diimide-Based Semiconductors: High Mobility n-Type Semiconductor for Organic Thin-Film Transistors. Chem. Mater. 2008, 20 (24), 7486−7491. (457) Chen, W.; Zhang, J.; Long, G.; Liu, Y.; Zhang, Q. From Nondetectable to Decent: Replacement of Oxygen with Sulfur in Naphthalene Diimide Boosts Electron Transport in Organic ThinFilm Transistors (OTFT). J. Mater. Chem. C 2015, 3 (31), 8219− 8224. (458) Kozycz, L. M.; Guo, C.; Manion, J. G.; Tilley, A. J.; Lough, A. J.; Li, Y.; Seferos, D. S. Enhanced Electron Mobility in Crystalline Thionated Naphthalene Diimides. J. Mater. Chem. C 2015, 3 (43), 11505−11515. (459) Nowak, E. M.; Sanetra, J.; Grucela, M.; Schab-Balcerzak, E. Azomethine Naphthalene Diimides as Component of Active Layers in Bulk Heterojunction Solar Cells. Mater. Lett. 2015, 157, 93−98. (460) Ke, H.; Weng, L.-J.; Chen, S.-Y.; Chen, J.-Z.; Lin, M.-J. Naphthalene Diimide Cocrystals: A Facile Approach to Tune the Optical Properties. Dyes Pigm. 2015, 113, 318−324. (461) Katz, H. E.; Johnson, J.; Lovinger, A. J.; Li, W. Naphthalenetetracarboxylic Diimide-Based n-Channel Transistor Semiconductors: Structural Variation and Thiol-Enhanced Gold Contacts. J. Am. Chem. Soc. 2000, 122 (32), 7787−7792. 11792

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Transistor Semiconductors. Chem. Commun. 2012, 48 (9), 1254− 1256. (478) Zhao, Z.; Zhang, F.; Zhang, X.; Yang, X.; Li, H.; Gao, X.; Di, C.-a.; Zhu, D. 1,2,5,6-Naphthalenediimide Based Donor−Acceptor Copolymers Designed from Isomer Chemistry for Organic Semiconducting Materials. Macromolecules 2013, 46 (19), 7705−7714. (479) Karpov, Y.; Maiti, J.; Tkachov, R.; Beryozkina, T.; Bakulev, V.; Liu, W.; Komber, H.; Lappan, U.; Al-Hussein, M.; Stamm, M.; Voit, B.; Kiriy, A. Copolymerization of Zinc-Activated Isoindigo- and Naphthalene-Diimide Based Monomers: an Efficient Route to Low Bandgap π-Conjugated Random Copolymers with Tunable Properties. Polym. Chem. 2016, 7 (15), 2691−2697. (480) Liu, J.-J.; Wang, Y.; Hong, Y.-J.; Lin, M.-J.; Huang, C.-C.; Dai, W.-X. A Photochromic Naphthalene Diimide Coordination Network Sensitized by Polyoxometalates. Dalton Trans. 2014, 43 (48), 17908− 17911. (481) Hu, J.-Y.; Nakano, M.; Osaka, I.; Takimiya, K. Naphthodithiophenediimide (NDTI)-Based Triads for High-Performance AirStable, Solution-Processed Ambipolar Organic Field-Effect Transistors. J. Mater. Chem. C 2015, 3 (17), 4244−4249. (482) Zhang, F.; Hu, Y.; Schuettfort, T.; Di, C.-a.; Gao, X.; McNeill, C. R.; Thomsen, L.; Mannsfeld, S. C. B.; Yuan, W.; Sirringhaus, H.; Zhu, D. Critical Role of Alkyl Chain Branching of Organic Semiconductors in Enabling Solution-Processed N-Channel Organic Thin-Film Transistors with Mobility of up to 3.50 cm2 V−1 s−1. J. Am. Chem. Soc. 2013, 135 (6), 2338−2349. (483) Gao, X.; Di, C.-a.; Hu, Y.; Yang, X.; Fan, H.; Zhang, F.; Liu, Y.; Li, H.; Zhu, D. Core-Expanded Naphthalene Diimides Fused with 2(1,3-Dithiol-2-Ylidene)Malonitrile Groups for High-Performance, Ambient-Stable, Solution-Processed n-Channel Organic Thin Film Transistors. J. Am. Chem. Soc. 2010, 132 (11), 3697−3699. (484) Zhang, F.; Zang, Y.; Huang, D.; Di, C.-a.; Gao, X.; Sirringhaus, H.; Zhu, D. Modulated Thermoelectric Properties of Organic Semiconductors Using Field-Effect Transistors. Adv. Funct. Mater. 2015, 25 (20), 3004−3012. (485) Gann, E.; Gao, X.; Di, C.-a.; McNeill, C. R. Phase Transitions and Anisotropic Thermal Expansion in High Mobility Core-expanded Naphthalene Diimide Thin Film Transistors. Adv. Funct. Mater. 2014, 24 (45), 7211−7220. (486) Hu, Y.; Gao, X.; Di, C.-a.; Yang, X.; Zhang, F.; Liu, Y.; Li, H.; Zhu, D. Core-Expanded Naphthalene Diimides Fused with Sulfur Heterocycles and End-Capped with Electron-Withdrawing Groups for Air-Stable Solution-Processed n-Channel Organic Thin Film Transistors. Chem. Mater. 2011, 23 (5), 1204−1215. (487) Liu, X.; Liu, Y.; Zheng, Y. Charge Transfer Mobility of Naphthodithiophenediimide Derivative: Normal-Mode and Bond Length Relaxation Analysis. Chem. Phys. Lett. 2016, 645, 92−96. (488) Suraru, S.-L.; Zschieschang, U.; Klauk, H.; Würthner, F. A core-extended naphthalene diimide as a p-channel semiconductor. Chem. Commun. 2011, 47 (41), 11504−11506. (489) Tan, L.; Guo, Y.; Zhang, G.; Yang, Y.; Zhang, D.; Yu, G.; Xu, W.; Liu, Y. New Air-Stable solution-Processed Organic n-Type Semiconductors Based on Sulfur-Rich Core-Expanded Naphthalene Diimides. J. Mater. Chem. 2011, 21 (44), 18042−18048. (490) Chang, J.; Ye, Q.; Huang, K.-W.; Zhang, J.; Chen, Z.-K.; Wu, J.; Chi, C. Stepwise Cyanation of Naphthalene Diimide for n-Channel Field-Effect Transistors. Org. Lett. 2012, 14 (12), 2964−2967. (491) Ortiz, R. P.; Herrera, H.; Blanco, R.; Huang, H.; Facchetti, A.; Marks, T. J.; Zheng, Y.; Segura, J. L. Organic n-Channel Field-Effect Transistors Based on Arylenediimide-Thiophene Derivatives. J. Am. Chem. Soc. 2010, 132 (24), 8440−8452. (492) Wang, Z.; Zhao, J.; Dong, H.; Qiu, G.; Zhang, Q.; Hu, W. An Asymmetric Naphthalimide Derivative for n-Channel Organic FieldEffect Transistors. Phys. Chem. Chem. Phys. 2015, 17 (40), 26519− 26524. (493) Jiang, J.-J.; Yan, C.; Pan, M.; Wang, Z.; Deng, H.-Y.; He, J.-R.; Yang, Q.-Y.; Fu, L.; Xu, X.-F.; Su, C.-Y. Structural Conformation and Optical and Electrochemical Properties of Imidazolyl-Substituted

(462) Sun, J.; Devine, R.; Dhar, B. M.; Jung, B. J.; See, K. C.; Katz, H. E. Improved Morphology and Performance from Surface Treatments of Naphthalenetetracarboxylic Diimide Bottom Contact Field-Effect Transistors. ACS Appl. Mater. Interfaces 2009, 1 (8), 1763−1769. (463) Stolte, M.; Gsanger, M.; Hofmockel, R.; Suraru, S.-L.; Würthner, F. Improved Ambient Operation of n-Channel Organic Transistors of Solution-Sheared Naphthalene Diimide Under Bias Stress. Phys. Chem. Chem. Phys. 2012, 14 (41), 14181−14185. (464) Geng, Y.; Wu, S.-X.; Li, H.-B.; Tang, X.-D.; Wu, Y.; Su, Z.-M.; Liao, Y. A Theoretical Discussion on the Relationships Among Molecular Packings, Intermolecular Interactions, and Electron Transport Properties for Naphthalene Tetracarboxylic Diimide Derivatives. J. Mater. Chem. 2011, 21 (39), 15558−15566. (465) See, K. C.; Landis, C.; Sarjeant, A.; Katz, H. E. Easily Synthesized Naphthalene Tetracarboxylic Diimide Semiconductors with High Electron Mobility in Air. Chem. Mater. 2008, 20 (11), 3609−3616. (466) Jung, B. J.; Sun, J.; Lee, T.; Sarjeant, A.; Katz, H. E. LowTemperature-Processible, Transparent, and Air-Operable n-Channel Fluorinated Phenylethylated Naphthalenetetracarboxylic Diimide Semiconductors Applied to Flexible Transistors. Chem. Mater. 2009, 21 (1), 94−101. (467) Lee, Y.-L.; Hsu, H.-L.; Chen, S.-Y.; Yew, T.-R. SolutionProcessed Naphthalene Diimide Derivatives as n-Type Semiconductor Materials. J. Phys. Chem. C 2008, 112 (5), 1694−1699. (468) Purdum, G. E.; Yao, N.; Woll, A.; Gessner, T.; Weitz, R. T.; Loo, Y.-L. Understanding Polymorph Transformations in CoreChlorinated Naphthalene Diimides and their Impact on Thin-Film Transistor Performance. Adv. Funct. Mater. 2016, 26 (14), 2357−2364. (469) Liu, L.; Ren, Z.; Xiao, C.; He, B.; Dong, H.; Yan, S.; Hu, W.; Wang, Z. Epitaxially-Crystallized Oriented Naphthalene Bis(dicarboximide) Morphology for Significant Performance Improvement of Electron-Transporting Thin-Film Transistors. Chem. Commun. 2016, 52 (27), 4902−4905. (470) Yuan, Z.; Ma, Y.; Geßner, T.; Li, M.; Chen, L.; Eustachi, M.; Weitz, R. T.; Li, C.; Müllen, K. Core-Fluorinated Naphthalene Diimides: Synthesis, Characterization, and Application in n-Type Organic Field-Effect Transistors. Org. Lett. 2016, 18 (3), 456−459. (471) Maniam, S.; Cox, R. P.; Langford, S. J.; Bell, T. D. M. Unexpected Photoluminescence of Fluorinated Naphthalene Diimides. Chem. - Eur. J. 2015, 21 (10), 4133−4140. (472) Zhang, X.; Wang, Z.; Chen, S.; Zhao, Z.; Yuan, W.; Wang, H.; Gao, X. Tuning the Charge Transport Property of Naphthalene Diimide Derivatives by Changing the Substituted Position of Fluorine Atom on Molecular Backbone. Chin. J. Chem. 2014, 32 (10), 1057− 1064. (473) He, T.; Stolte, M.; Burschka, C.; Hansen, N. H.; Musiol, T.; Kälblein, D.; Pflaum, J.; Tao, X.; Brill, J.; Würthner, F. Single-crystal Field-effect Transistors of new Cl2-NDI Polymorph Processed by Sublimation in Air. Nat. Commun. 2015, 6, 5954. (474) Hansen, M. H.; May, F.; Kälblein, D.; Schmeiler, T.; Lennartz, C.; Steeger, A.; Burschka, C.; Stolte, M.; Würthner, F.; Brill, J.; Pflaum, J. Anisotropic Electron Mobility Studies on Cl2-NDI Single Crystals and the Role of Static and Dynamic Lattice Deformations upon Temperature Variation. cond-mat.mtrl-sci 2015, e-Print Archive (arXiv:1501.01856), 1−19. (475) Jung, J. W.; Jo, J. W.; Chueh, C.-C.; Liu, F.; Jo, W. H.; Russell, T. P.; Jen, A. K. Y. Fluoro-Substituted n-Type Conjugated Polymers for Additive-Free All-Polymer Bulk Heterojunction Solar Cells with High Power Conversion Efficiency of 6.71%. Adv. Mater. 2015, 27 (21), 3310−3317. (476) Luzio, A.; Fazzi, D.; Nübling, F.; Matsidik, R.; Straub, A.; Komber, H.; Giussani, E.; Watkins, S. E.; Barbatti, M.; Thiel, W.; Gann, E.; Thomsen, L.; McNeill, C. R.; Caironi, M.; Sommer, M. Structure−Function Relationships of High-Electron Mobility Naphthalene Diimide Copolymers Prepared Via Direct Arylation. Chem. Mater. 2014, 26 (21), 6233−6240. (477) Chen, S.-c.; Zhang, Q.; Zheng, Q.; Tang, C.; Lu, C.-Z. AngularShaped Naphthalene Tetracarboxylic Diimides for n-Channel Organic 11793

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

Naphthalenediimide and Its HgII, CdII, and CuII Halide Complexes. Eur. J. Inorg. Chem. 2012, 2012 (8), 1171−1179. (494) Takahashi, E.; Takaya, H.; Naota, T. Dynamic Vapochromic Behaviors of Organic Crystals Based on the Open−Close Motions of S-Shaped Donor−Acceptor Folding Units. Chem. - Eur. J. 2010, 16 (16), 4793−4802. (495) Zheng, J.; Qiao, W.; Wan, X.; Gao, J. P.; Wang, Z. Y. NearInfrared Electrochromic and Chiroptical Switching Materials: Design, Synthesis, and Characterization of Chiral Organogels Containing Stacked Naphthalene Diimide Chromophores. Chem. Mater. 2008, 20 (19), 6163−6168. (496) Kondratenko, M.; Moiseev, A. G.; Perepichka, D. F. New Stable Donor-Acceptor Dyads for Molecular Electronics. J. Mater. Chem. 2011, 21 (5), 1470−1478. (497) Lin, C.-C.; Velusamy, M.; Chou, H.-H.; Lin, J. T.; Chou, P.-T. Synthesis and Characterization of Naphthalene Diimide (NDI)-Based Near Infrared Chromophores with Two-Photon Absorbing Properties. Tetrahedron 2010, 66 (45), 8629−8634. (498) Bobe, S. R.; Gupta, A.; Rananaware, A.; Bilic, A.; Xiang, W.; Li, J.; Bhosale, S. V.; Bhosale, S. V.; Evans, R. A. Insertion of a Naphthalenediimide Unit in a Metal-free Donor−Acceptor Organic Sensitizer for Efficiency Enhancement of a Dye-sensitized Solar Cell. Dyes Pigm. 2016, 134, 83. (499) Fernando, R.; Etheridge, F.; Muller, E.; Sauve, G. Tuning the Optical and Electrochemical Properties of Core-substituted Naphthalenediimides with Styryl Imide Substituent. New J. Chem. 2015, 39 (4), 2506−2514. (500) Kumar, S.; Ajayakumar, M. R.; Hundal, G.; Mukhopadhyay, P. Extraordinary Stability of Naphthalenediimide Radical Ion and Its Ultra-Electron-Deficient Precursor: Strategic Role of the Phosphonium Group. J. Am. Chem. Soc. 2014, 136 (34), 12004−12010. (501) Li, Y.; Zhang, G.; Zhang, W.; Wang, J.; Chen, X.; Liu, Z.; Yan, Y.; Zhao, Y.; Zhang, D. Arylacetylene-Substituted Naphthalene Diimides with Dual Functions: Optical Waveguides and n-Type Semiconductors. Chem. - Asian J. 2014, 9 (11), 3207−3214. (502) Polander, L. E.; Barlow, S.; Seifried, B. M.; Marder, S. R. A 2,6Diformylnaphthalene-1,8:4,5-bis(dicarboximide): Synthesis and Knoevenagel Condensation with Malononitrile. J. Org. Chem. 2012, 77 (20), 9426−9428. (503) Etheridge, F. S.; Fernando, R.; Golen, J. A.; Rheingold, A. L.; Sauve, G. Tuning the Optoelectronic Properties of Core-substituted Naphthalene Diimides by the Selective Conversion of Imides to Monothioimides. RSC Adv. 2015, 5 (58), 46534−46539. (504) Gabutti, S.; Schaffner, S.; Neuburger, M.; Fischer, M.; Schafer, G.; Mayor, M. Planar Chiral Asymmetric Naphthalenediimide Cyclophanes: Synthesis, Characterization and Tunable FRET Properties. Org. Biomol. Chem. 2009, 7 (16), 3222−3229. (505) Liu, Y.; Zhang, L.; Lee, H.; Wang, H.-W.; Santala, A.; Liu, F.; Diao, Y.; Briseno, A. L.; Russell, T. P. NDI-Based Small Molecule as Promising Nonfullerene Acceptor for Solution-Processed Organic Photovoltaics. Adv. Energy Mater. 2015, 5 (12), 1500195. (506) Ahmed, E.; Ren, G.; Kim, F. S.; Hollenbeck, E. C.; Jenekhe, S. A. Design of New Electron Acceptor Materials for Organic Photovoltaics: Synthesis, Electron Transport, Photophysics, and Photovoltaic Properties of Oligothiophene-Functionalized Naphthalene Diimides. Chem. Mater. 2011, 23 (20), 4563−4577. (507) Hwang, D. K.; Dasari, R. R.; Fenoll, M.; Alain-Rizzo, V.; Dindar, A.; Shim, J. W.; Deb, N.; Fuentes-Hernandez, C.; Barlow, S.; Bucknall, D. G.; Audebert, P.; Marder, S. R.; Kippelen, B. Stable Solution-Processed Molecular n-Channel Organic Field-Effect Transistors. Adv. Mater. 2012, 24 (32), 4445−4450. (508) Polander, L. E.; Romanov, A. S.; Barlow, S.; Hwang, D. K.; Kippelen, B.; Timofeeva, T. V.; Marder, S. R. Stannyl Derivatives of Naphthalene Diimides and Their Use in Oligomer Synthesis. Org. Lett. 2012, 14 (3), 918−921. (509) Takai, A.; Sakamaki, D.; Seki, S.; Matsushita, Y.; Takeuchi, M. Ferrocene-Substituted Naphthalenediimide with Broad Absorption and Electron-Transport Properties in the Segregated-Stack Structure. Chem. - Eur. J. 2016, 22 (22), 7385−7388.

(510) Chen, Z.; Zheng, Y.; Yan, H.; Facchetti, A. Naphthalenedicarboximide- vs Perylenedicarboximide-Based Copolymers. Synthesis and Semiconducting Properties in Bottom-Gate N-Channel Organic Transistors. J. Am. Chem. Soc. 2009, 131 (1), 8−9. (511) Lee, C.; Kang, H.; Lee, W.; Kim, T.; Kim, K.-H.; Woo, H. Y.; Wang, C.; Kim, B. J. High-Performance All-Polymer Solar Cells via Side-Chain Engineering of the Polymer Acceptor: The Importance of the Polymer Packing Structure and the Nanoscale Blend Morphology. Adv. Mater. 2015, 27 (15), 2466−2471. (512) Choi, J.; Kim, K.-H.; Yu, H.; Lee, C.; Kang, H.; Song, I.; Kim, Y.; Oh, J. H.; Kim, B. J. Importance of Electron Transport Ability in Naphthalene Diimide-Based Polymer Acceptors for High-Performance, Additive-Free, All-Polymer Solar Cells. Chem. Mater. 2015, 27 (15), 5230−5237. (513) Xia, Y.; Musumeci, C.; Bergqvist, J.; Ma, W.; Gao, F.; Tang, Z.; Bai, S.; Jin, Y.; Zhu, C.; Kroon, R.; Wang, C.; Andersson, M. R.; Hou, L.; Inganas, O.; Wang, E. Inverted All-Polymer Solar Cells Based on a Quinoxaline-Thiophene/Naphthalene-Diimide Polymer Blend Improved by Annealing. J. Mater. Chem. A 2016, 4, 3835. (514) Kurosawa, T.; Chiu, Y.-C.; Zhou, Y.; Gu, X.; Chen, W.-C.; Bao, Z. Impact of Polystyrene Oligomer Side Chains on Naphthalene Diimide−Bithiophene Polymers as n-Type Semiconductors for Organic Field-Effect Transistors. Adv. Funct. Mater. 2016, 26 (8), 1261−1270. (515) An, Y.; Long, D. X.; Kim, Y.; Noh, Y.-Y.; Yang, C. Improved Electron Transport Properties of n-Type Naphthalenediimide Polymers Through Refined Molecular Ordering and Orientation Induced by Processing Solvents. Phys. Chem. Chem. Phys. 2016, 18 (18), 12486−12493. (516) Xiao, B.; Ding, G.; Tan, Z. a.; Zhou, E. A Comparison of nType Copolymers Based on Cyclopentadithiophene and Naphthalene Diimide/Perylene Diimides for All-Polymer Solar Cell Applications. Polym. Chem. 2015, 6 (43), 7594−7602. (517) Geng, Y.; Huang, J.; Tajima, K.; Zeng, Q.; Zhou, E. A Low Band Gap n-Type Polymer Based on Dithienosilole and Naphthalene Diimide for All-Polymer Solar Cells Application. Polymer 2015, 63, 164−169. (518) Wu, Z.; Sun, C.; Dong, S.; Jiang, X.-F.; Wu, S.; Wu, H.; Yip, H.L.; Huang, F.; Cao, Y. n-Type Water/Alcohol-Soluble Naphthalene Diimide-Based Conjugated Polymers for High-Performance Polymer Solar Cells. J. Am. Chem. Soc. 2016, 138 (6), 2004−2013. (519) Sun, C.; Wu, Z.; Yip, H.-L.; Zhang, H.; Jiang, X.-F.; Xue, Q.; Hu, Z.; Hu, Z.; Shen, Y.; Wang, M.; Huang, F.; Cao, Y. Solar Cells: Amino-Functionalized Conjugated Polymer as an Efficient Electron Transport Layer for High-Performance Planar-Heterojunction Perovskite Solar Cells. Adv. Energy Mater. 2016, 6 (5), n/a. (520) Nam, S. Y.; Choi, E. Y.; Song, C. E.; Lee, C.; Jung, I. H.; Yoon, S. C. Naphthalene-Diimide-Incorporated Conjugated Polyelectrolyte Interfacial Modifier for the Efficient Inverted-Type Polymer Solar Cells. J. Inf. Disp. 2016, 17 (1), 17−24. (521) Matsidik, R.; Komber, H.; Luzio, A.; Caironi, M.; Sommer, M. Defect-free Naphthalene Diimide Bithiophene Copolymers with Controlled Molar Mass and High Performance via Direct Arylation Polycondensation. J. Am. Chem. Soc. 2015, 137 (20), 6705−6711. (522) Senkovskyy, V.; Tkachov, R.; Komber, H.; Sommer, M.; Heuken, M.; Voit, B.; Huck, W. T. S.; Kataev, V.; Petr, A.; Kiriy, A. Chain-Growth Polymerization of Unusual Anion-Radical Monomers Based on Naphthalene Diimide: A New Route to Well-Defined nType Conjugated Copolymers. J. Am. Chem. Soc. 2011, 133 (49), 19966−19970. (523) Durban, M. M.; Kazarinoff, P. D.; Luscombe, C. K. Synthesis and Characterization of Thiophene-Containing Naphthalene Diimide n-Type Copolymers for OFET Applications. Macromolecules 2010, 43 (15), 6348−6352. (524) Zhou, W.; Wen, Y.; Ma, L.; Liu, Y.; Zhan, X. Conjugated Polymers of Rylene Diimide and Phenothiazine for n-Channel Organic Field-Effect Transistors. Macromolecules 2012, 45 (10), 4115−4121. 11794

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

with Improved Open-Circuit Voltage and Morphology. RSC Adv. 2015, 5 (112), 92151−92158. (542) Murari, N. M.; Hwang, Y.-J.; Kim, F. S.; Jenekhe, S. A. Organic Nonvolatile Memory Devices Utilizing Intrinsic Charge-Trapping Phenomena in an n-Type Polymer Semiconductor. Org. Electron. 2016, 31, 104−110. (543) Hwang, Y.-J.; Courtright, B. A. E.; Jenekhe, S. A. Ternary Blend All-Polymer Solar Cells: Enhanced Performance and Evidence of Parallel-like Bulk Heterojunction Mechanism. MRS Commun. 2015, 5 (02), 229−234. (544) Goto, E.; Ando, S.; Ueda, M.; Higashihara, T. Nonstoichiometric Stille Coupling Polycondensation for Synthesizing Naphthalene-Diimide-Based π-Conjugated Polymers. ACS Macro Lett. 2015, 4 (9), 1004−1007. (545) Fukuta, S.; Wu, H.-C.; Koganezawa, T.; Isshiki, Y.; Ueda, M.; Chen, W.-C.; Higashihara, T. Synthesis and FET Characterization of Novel Ambipolar and Low-Bandgap Naphthalene-Diimide-based Semiconducting Polymers. J. Polym. Sci., Part A: Polym. Chem. 2016, 54 (3), 359−367. (546) Goto, E.; Mori, H.; Ueda, M.; Higashihara, T. Controlled Polymerization of Electron-deficient Naphthalene-diimide Containing Monomer by Negishi-type Catalyst-transfer Polymerization. J. Photopolym. Sci. Technol. 2015, 28 (2), 279−283. (547) Schroot, R.; Schlotthauer, T.; Schubert, U. S.; Jäger, M. Modular Assembly of Poly(naphthalene diimide) and Ru(II) Dyes for an Efficient Light-Induced Charge Separation in Hierarchically Controlled Polymer Architectures. Macromolecules 2016, 49 (6), 2112−2123. (548) Prentice, G. M.; Emmett, L. M.; Zhu, H.; Kociok-Köhn, G.; Pantoş, G. D. Thermally Stable Recyclable Naphthalenediimide− Siloxane Polymers. Supramol. Chem. 2016, 28 (1−2), 161−167. (549) Zhang, L.; Xu, H.; Ma, X.; Shi, Z.; Yin, J.; Jiang, X. Toward Multifunctional Polymer Hybrid through Tunable Charge Transfer Interaction of Anthracene/Naphthalenediimide. Adv. Mater. Interfaces 2016, 3, 1600224. (550) Zhao, Z.; Wang, Z.; Ge, C.; Zhang, X.; Yang, X.; Gao, X. Incorporation of Benzothiadiazole into the Backbone of 1,2,5,6Naphthalenediimide Based Copolymers, Enabling Much Improved Film Crystallinity and Charge Carrier Mobility. Polym. Chem. 2016, 7 (3), 573−579. (551) Li, Z.; Feng, K.; Liu, J.; Mei, J.; Li, Y.; Peng, Q. Large BandGap Copolymers Based on a 1,2,5,6-Naphthalenediimide Unit for Polymer Solar Cells with High Open Circuit Voltages and Power Conversion Efficiencies. J. Mater. Chem. A 2016, 4 (19), 7372−7381. (552) DeBlase, C. R.; Hernández-Burgos, K.; Rotter, J. M.; Fortman, D. J.; dos S. Abreu, D.; Timm, R. A.; Diógenes, I. C. N.; Kubota, L. T.; Abruña, H. D.; Dichtel, W. R. Cation-Dependent Stabilization of Electrogenerated Naphthalene Diimide Dianions in Porous Polymer Thin Films and Their Application to Electrical Energy Storage. Angew. Chem., Int. Ed. 2015, 54 (45), 13225−13229. (553) Szumilo, M. M.; Gann, E. H.; McNeill, C. R.; Lemaur, V.; Oliver, Y.; Thomsen, L.; Vaynzof, Y.; Sommer, M.; Sirringhaus, H. Structure Influence on Charge Transport in NaphthalenediimideThiophene Copolymers. Chem. Mater. 2014, 26 (23), 6796−6804. (554) Liu, Y.; Page, Z. A.; Russell, T. P.; Emrick, T. Finely Tuned Polymer Interlayers Enhance Solar Cell Efficiency. Angew. Chem., Int. Ed. 2015, 54 (39), 11485−11489. (555) Zeigler, D. F.; Candelaria, S. L.; Mazzio, K. A.; Martin, T. R.; Uchaker, E.; Suraru, S.-L.; Kang, L. J.; Cao, G.; Luscombe, C. K. NType Hyperbranched Polymers for Supercapacitor Cathodes with Variable Porosity and Excellent Electrochemical Stability. Macromolecules 2015, 48 (15), 5196−5203. (556) Qin, Y.; Li, X.; Sun, W.; Luo, X.; Li, M.; Tang, X.; Jin, X.; Xie, Y.; Ouyang, X.; Li, Q. Small Bandgap Naphthalene Diimide Copolymers for Efficient Inorganic-Organic Hybrid Solar Cells. RSC Adv. 2015, 5 (3), 2147−2154. (557) Wang, P.; Li, H.; Gu, C.; Dong, H.; Xu, Z.; Fu, H. Air-Stable Ambipolar Organic Field-Effect Transistors Based on Naphthalene-

(525) Popere, B. C.; Della Pelle, A. M.; Thayumanavan, S. BODIPYBased Donor−Acceptor π-Conjugated Alternating Copolymers. Macromolecules 2011, 44 (12), 4767−4776. (526) Alvey, P. M.; Iverson, B. L. Reactions of Brominated Naphthalene Diimide with Bis(tributylstannyl)acetylene: A Simple Approach for Conjugated Polymers and Versatile Coupling Intermediates. Org. Lett. 2012, 14 (11), 2706−2709. (527) Kim, Y.; Cho, H.-H.; Kim, T.; Liao, K.; Kim, B. J. Terpolymer Approach for Controlling the Crystalline Behavior of Naphthalene Diimide-Based Polymer Acceptors and Enhancing the Performance of All-Polymer Solar Cells. Polym. J. 2016, 48 (4), 517−524. (528) Xue, L.; Yang, Y.; Zhang, Z.-G.; Dong, X.; Gao, L.; Bin, H.; Zhang, J.; Yang, Y.; Li, Y. Indacenodithienothiophene-Naphthalene Diimide Copolymer as an Acceptor for All-Polymer Solar Cells. J. Mater. Chem. A 2016, 4 (16), 5810−5816. (529) Kim, Y.; Long, D. X.; Lee, J.; Kim, G.; Shin, T. J.; Nam, K.-W.; Noh, Y.-Y.; Yang, C. A Balanced Face-On to Edge-On Texture Ratio in Naphthalene Diimide-Based Polymers with Hybrid Siloxane Chains Directs Highly Efficient Electron Transport. Macromolecules 2015, 48 (15), 5179−5187. (530) Shao, J.; Wang, G.; Wang, K.; Yang, C.; Wang, M. Direct Arylation Polycondensation for Efficient Synthesis of Narrow-Bandgap Alternating D-A Copolymers Consisting of Naphthalene Diimide as an Acceptor. Polym. Chem. 2015, 6 (38), 6836−6844. (531) Dai, S.; Cheng, P.; Lin, Y.; Wang, Y.; Ma, L.; Ling, Q.; Zhan, X. Perylene and Naphthalene Diimide Polymers for All-Polymer Solar Cells: a Comparative Study of Chemical Copolymerization and Physical Blend. Polym. Chem. 2015, 6 (29), 5254−5263. (532) Zhao, K.; Ye, L.; Zhao, W.; Zhang, S.; Yao, H.; Xu, B.; Sun, M.; Hou, J. Enhanced Efficiency of Polymer Photovoltaic Cells via the Incorporation of a Water-Soluble Naphthalene Diimide Derivative as a Cathode Interlayer. J. Mater. Chem. C 2015, 3 (37), 9565−9571. (533) Ye, L.; Jiao, X.; Zhang, H.; Li, S.; Yao, H.; Ade, H.; Hou, J. 2DConjugated Benzodithiophene-Based Polymer Acceptor: Design, Synthesis, Nanomorphology, and Photovoltaic Performance. Macromolecules 2015, 48 (19), 7156−7163. (534) Li, C.; Zhang, A.; Wang, Z.; Liu, F.; Zhou, Y.; Russell, T. P.; Li, Y.; Li, W. All Polymer Solar Cells with Diketopyrrolopyrrole-Polymers as Electron Donor and a Naphthalenediimide-Polymer as Electron Acceptor. RSC Adv. 2016, 6 (42), 35677−35683. (535) Lee, W.; Lee, C.; Yu, H.; Kim, D.-J.; Wang, C.; Woo, H. Y.; Oh, J. H.; Kim, B. J. Side Chain Optimization of Naphthalenediimide− Bithiophene-Based Polymers to Enhance the Electron Mobility and the Performance in All-Polymer Solar Cells. Adv. Funct. Mater. 2016, 26 (10), 1543−1553. (536) Anton, A. M.; Steyrleuthner, R.; Kossack, W.; Neher, D.; Kremer, F. Spatial Orientation and Order of Structure-Defining Subunits in Thin Films of a High Mobility n-Type Copolymer. Macromolecules 2016, 49 (5), 1798−1806. (537) Hwang, Y.-J.; Earmme, T.; Courtright, B. A. E.; Eberle, F. N.; Jenekhe, S. A. n-Type Semiconducting Naphthalene Diimide-Perylene Diimide Copolymers: Controlling Crystallinity, Blend Morphology, and Compatibility Toward High-Performance All-Polymer Solar Cells. J. Am. Chem. Soc. 2015, 137 (13), 4424−4434. (538) Kozycz, L. M.; Gao, D.; Tilley, A. J.; Seferos, D. S. One Donor−two Acceptor (D−A1)-(D−A2) Random Terpolymers Containing Perylene Diimide, Naphthalene Diimide, and Carbazole Units. J. Polym. Sci., Part A: Polym. Chem. 2014, 52 (23), 3337−3345. (539) Earmme, T.; Hwang, Y.-J.; Subramaniyan, S.; Jenekhe, S. A. All-Polymer Bulk Heterojuction Solar Cells with 4.8% Efficiency Achieved by Solution Processing from a Co-Solvent. Adv. Mater. 2014, 26 (35), 6080−6085. (540) Mori, D.; Benten, H.; Okada, I.; Ohkita, H.; Ito, S. Highly Efficient Charge-Carrier Generation and Collection in Polymer/ Polymer Blend Solar Cells with a Power Conversion Efficiency of 5.7%. Energy Environ. Sci. 2014, 7 (9), 2939−2943. (541) Liu, F.; Li, H.; Wu, Y.; Gu, C.; Fu, H. Naphthalene Diimide and Benzothiadiazole Copolymer Acceptor for All-polymer Solar Cells 11795

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796

Chemical Reviews

Review

diimide-Diketopyrrolopyrrole Copolymers. RSC Adv. 2015, 5 (25), 19520−19527. (558) Schottel, B. L.; Chifotides, H. T.; Dunbar, K. R. Anion-π Interactions. Chem. Soc. Rev. 2008, 37 (1), 68−83. (559) Chifotides, H. T.; Dunbar, K. R. Anion−π Interactions in Supramolecular Architectures. Acc. Chem. Res. 2013, 46 (4), 894−906. (560) Caltagirone, C.; Gale, P. A. Anion Receptor Chemistry: Highlights from 2007. Chem. Soc. Rev. 2009, 38 (2), 520−563. (561) Vicic, D. A.; Odom, D. T.; Núñez, M. E.; Gianolio, D. A.; McLaughlin, L. W.; Barton, J. K. Oxidative Repair of a Thymine Dimer in DNA from a Distance by a Covalently Linked Organic Intercalator. J. Am. Chem. Soc. 2000, 122 (36), 8603−8611. (562) Kim, M. B.; Dixon, D. W. Hydrolysis of Aliphatic Naphthalene Diimides: Effect of Charge Placement in the Side Chains. J. Phys. Org. Chem. 2008, 21 (9), 731−737. (563) Fujisawa, K.; Humbert-Droz, M.; Letrun, R.; Vauthey, E.; Wesolowski, T. A.; Sakai, N.; Matile, S. Ion Pair−π Interactions. J. Am. Chem. Soc. 2015, 137 (34), 11047−11056. (564) Fujisawa, K.; Beuchat, C.; Humbert-Droz, M.; Wilson, A.; Wesolowski, T. A.; Mareda, J.; Sakai, N.; Matile, S. Anion−π and Cation−π Interactions on the Same Surface. Angew. Chem., Int. Ed. 2014, 53 (42), 11266−11269. (565) Akamatsu, M.; Matile, S. Expanded Chiral Surfaces for Asymmetric Anion−π Catalysis. Synlett 2016, 27 (07), 1041−1046. (566) Zhao, Y.; Beuchat, C.; Domoto, Y.; Gajewy, J.; Wilson, A.; Mareda, J.; Sakai, N.; Matile, S. Anion−π Catalysis. J. Am. Chem. Soc. 2014, 136 (5), 2101−2111. (567) Zhao, Y.; Cotelle, Y.; Avestro, A.-J.; Sakai, N.; Matile, S. Asymmetric Anion-π Catalysis: Enamine Addition to Nitroolefins on π-Acidic Surfaces. J. Am. Chem. Soc. 2015, 137 (36), 11582−11585. (568) Zhao, Y.; Huang, G.; Besnard, C.; Mareda, J.; Sakai, N.; Matile, S. Big, Strong, Neutral, Twisted, and Chiral π Acids. Chem. - Eur. J. 2015, 21 (16), 6202−6207. (569) Copeland, G. T.; Miller, S. J. Selection of Enantioselective Acyl Transfer Catalysts from a Pooled Peptide Library through a Fluorescence-Based Activity Assay: An Approach to Kinetic Resolution of Secondary Alcohols of Broad Structural Scope. J. Am. Chem. Soc. 2001, 123 (27), 6496−6502. (570) Lin, N.-T.; Vargas Jentzsch, A.; Guenee, L.; Neudorfl, J.-M.; Aziz, S.; Berkessel, A.; Orentas, E.; Sakai, N.; Matile, S. Enantioselective Self-Sorting on Planar, π-acidic Surfaces of Chiral Anion-π Transporters. Chem. Sci. 2012, 3 (4), 1121−1127. (571) Miros, F. N.; Zhao, Y.; Sargsyan, G.; Pupier, M.; Besnard, C.; Beuchat, C.; Mareda, J.; Sakai, N.; Matile, S. Enolate Stabilization by Anion−π Interactions: Deuterium Exchange in Malonate Dilactones on π-Acidic Surfaces. Chem. - Eur. J. 2016, 22, 2648−2657. (572) Zhao, Y.; Benz, S.; Sakai, N.; Matile, S. Selective Acceleration of Disfavored Enolate Addition Reactions by Anion-π Interactions. Chem. Sci. 2015, 6 (11), 6219−6223. (573) Zhao, Y.; Sakai, N.; Matile, S. Enolate Chemistry with Anion−π Interactions. Nat. Commun. 2014, 5, No. 3911, DOI: 10.1038/ncomms4911. (574) Cotelle, Y.; Benz, S.; Avestro, A.-J.; Ward, T. R.; Sakai, N.; Matile, S. Anion−π Catalysis of Enolate Chemistry: Rigidified Leonard Turns as a General Motif to Run Reactions on Aromatic Surfaces. Angew. Chem., Int. Ed. 2016, 55, 4275−4279.

11796

DOI: 10.1021/acs.chemrev.6b00160 Chem. Rev. 2016, 116, 11685−11796