Perylene Diimide Bearing Different Trialkyl Silyl Ethers: Impact of

Apr 30, 2018 - For example, PDI with one imide functionalized with a nonpolar alkyl chain and the other with a propylene oxide–ethylene oxide copoly...
0 downloads 5 Views 2MB Size
Subscriber access provided by Kaohsiung Medical University

Article

Perylene diimide bearing different trialkyl silyl ethers: Impact of asymmetric functionalization on self-assembly into nanostructures Rachael Matthews, Jordan Swisher, Kristin M. Hutchins, and Emily B. Pentzer Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.8b01543 • Publication Date (Web): 30 Apr 2018 Downloaded from http://pubs.acs.org on May 1, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Perylene diimide bearing different trialkyl silyl ethers: Impact of asymmetric functionalization on self-assembly into nanostructures Rachael Matthews,1 Jordan Swisher,2 Kristin M. Hutchins,3 Emily B. Pentzer1,* 1

Case Western Reserve University, Department of Chemistry, Cleveland, OH 44106 Now at University of Pittsburgh, Department of Chemistry, Pittsburgh, PA 15260 3 Texas Tech University, Department of Chemistry & Biochemistry, Lubbock, TX 79409 2

ABSTRACT: For over a decade, a great amount of research effort has focused on controlling the size and shape of organic small molecule crystals, as these parameters impact physical and optoelectronic properties. A thorough understanding of how functionalization impacts assembly, as well as guiding principles to control aggregation and self-assembly, are vital to producing novel organic nanostructures for electronic applications such as organic photovoltaics (OPVs). Herein, we study the influence of unsymmetrical functionalization of perylene diimide (PDI) on self-assembly. The guiding hypothesis of this work is that the identity of the pendant functionalities will impact the size, aspect ratio, and surface properties of the resulting assemblies. Twelve asymmetrically functionalized PDI molecules are reported, in which the length of the alkyl substituents at the imide position is varied, and include alcohol and silylated alcohol functionalities at the end of the alky chain. Morphologies of these self-assembled structures were characterized by scanning and transmission electronic microscopy, crystallinity was verified by powder X-ray diffraction, and the optoelectronic and thermal properties are also reported. Based on the functionality of the PDI molecules, different shaped assemblies are prepared, including high aspect ratio structures with widths ranging from 0.1-2.5 µm and lengths 1- 800 µm.

Introduction Perylene diimide (PDI) is a rylene dye used in industrial paints and is also a well-known organic n-type semiconductor.1–7 PDI exhibits favorable properties such as high photochemical stability, high molar absorption coefficient, low cost, nearunity fluorescence quantum yields, and charge carrier mobility as high as 0.1 cm2/Vs in the liquid crystalline form.8 As such, PDI and its functionalized derivatives are attractive for use as active materials in a variety of (opto)electronic devices, including organic field-effect transistors (OFETs), dye lasers, and organic photovoltaics (OPVs).1,2,7–12 Typically, PDI derivatives with symmetric substitution at the imide positions have been studied, specifically the imide groups have been functionalized with a variety of alkyl groups. Such functionalization controls both the solubility and aggregation/self-assembly of the molecules and can impact device performance.2,8,11,13–15 Compared to functionalization of the aromatic rings of PDI at the bay positions, functionalization of the imides does not impact the electronic properties of the molecule, as both the HOMO and LUMO have nodes at the imide position.10,11,14,16 Controlling the self-assembly of organic molecules into ordered structures is a challenging feat, but will have positive impacts on diverse fields including biochemistry, engineering, and green chemistry, among others.8,17,18 The morphology of self-assembled PDI depends on the alkyl functionalization of the imide positions, as these substituents dictate the balance of molecule-molecule interactions by π-π stacking of aromatic systems and hydrophobic interactions of alkyl chains.8,13–16,19–21 Longer alkyl chains provide solubility and thus processability in common organic solvents, however the insulating nature of these alkyl substituents can lower device efficiency.10,14,22 Typically, self-assembly of PDI into

nanostructures is achieved by one of two methods: 1) a solvent evaporation method, in which solvent slowly evaporates from a concentrated solution of the molecule; or 2) a solvent diffusion process7 in which a bad solvent is added on top of a solution of PDI in a good solvent and self-assembly occurs at the interface of the two liquids. Symmetrically functionalized PDI molecules and their assemblies include: n-alkyl and branched substituents that result in nanowires;9 ethoxy substituents resulting in nanobelts;14,21,23 and swallowtail substituents that give spherical nanostructures.14 More exotic substituents have been used to access nanostructures of intricate design, including: glucopyranoside, polyhedral oligosilsesquioxane (POSS), and aniline that result in helical nanofibers,17 single crystalline nanobelts,24 and micellular shapes,25 respectively. Thus, for symmetrically substituted PDI molecules, the chemical nature of the imide functionalization plays a vital role in the structures formed.8,18 Alternatively, unsymmetrically functionalized PDI derivatives, in which each imide position is functionalized with a distinct functionality, have gained attention, though they are significantly more difficult to synthesize than their symmetrically functionalized counterparts.8,24,26 Most commonly, for asymmetrically functionalized PDI molecules one imide contains a solubilizing branched alkyl group27 and the other imide position contains a polar functionality. For example, PDI with one imide functionalized with a nonpolar alkyl chain and the other with a propylene oxide–ethylene oxide copolymer selfassembled into hollow nanotubes;24 other polar substituents and nanostructures include alkoxy substituents yielding nanocoils,28,29 poly-oxyethylene giving nanobelts,27 and fluoroalkylated groups forming nanoribbons.26 Expanding the chemical functionalities that can be used to prepare asymmetrically

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1. Overview of the work reported herein: asymmetrically functionalized perylene diimide (PDI) derivatives bearing different functionalities at the two imide positions to give different sized building blocks that self-assemble into nanostructures. functionalized PDIs and understanding their impact on selfassembly will facilitate the development of next generation materials for optoelectronic devices.13,30 Herein we present the synthesis and characterization of twelve different asymmetrically functionalized PDI small molecules and study their self-assembly using the solvent diffusion method. One imide position of the PDI is functionalized with a short alkyl chain for solubility and the other imide position is functionalized with an amino alcohol; the hydroxyl group of the amino alcohol is subsequently transformed to a trialkyl silyl ether using standard conditions to give different shaped molecular “building blocks” (Figure 1). We demonstrate that the identity of the amino alcohol and size of the trialkyl silyl group impact the shape and size of the crystals

Page 2 of 9

formed, but that the interplay of these two factors cannot be disentangled. The thermal properties of the materials are characterized by thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) and the self-assembled structures are characterized by UV-Vis and fluorescence spectroscopies, scanning electron microscopy (SEM), transmission electron microscopy (TEM) and powder X-ray diffraction (XRD). Regardless of the functionality of the imide substituents and size and shape of the self-assembled structures, all are crystalline, however they have distinct size and shape based on differences in the functionalities. Results and Discussion Synthesis and Self-assembly. An overview of the synthesis of the asymmetrically functionalized PDI derivatives is shown in Figure 2A, along with the amino alcohol, trialkyl silyl groups, and the naming scheme used (Figure 2B). Briefly, perylene dianhydride was converted to the diimide derivative using 3-aminopentane, then one of the imides underwent saponification by tandem treatment with strong base and strong acid to yield perylene monoimide monoanhydride (PMIMA). The anhydride of PMIMA was functionalized with an amino alcohol to give three different asymmetrically functionalized PDI molecules: PDI 1-OH, PDI 2-OH, and PDI 3-OH. In these molecules, the branched alkyl chain provides solubility in organic solvents as well as alkyl-alkyl interactions during selfassembly, and the pendant alcohol functionality on the other imide position provides a handle for functionalization to control the overall shape of the molecules. The alcohols of these molecules were functionalized with different trialkylsilyl chlorides to give the corresponding silyl ethers: trimethyl silyl (TMS), tertbutyldimethylsilyl (TBDMS), and triisopropylsilyl (TIPS). All PDI derivatives were characterized by 1H and 13C NMR (Figures S1-S15), FTIR (Figure S16), and MALDI-TOF (Figure S17).

Figure 2. A) Synthetic procedure for the preparation of asymmetrically functionalized PDI molecules; B) Asymmetric PDI moleACS Paragon Plus Environment cules and naming scheme used herein. i = imidazole, 125 °C, 2-amino pentane; ii = 1) KOH, tBuOH, 90 °C, 2) AcOH, 2N HCl; iii = imidazole, 125 °C, amino-alcohol; iv = imidazole, 125 °C, R3Si-Cl.

Page 3 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 3. Optoelectronic data. A) UV-vis absorption spectra of PDI A in chloroform (black), a suspension of the aggregates (red), and aggregates in the solid state deposited on glass (purple); B) Fluorescence emission spectra of PDI A solvated in chloroform (black) and aggregates in suspension (purple). To determine the impact of the molecular functionality on self-assembly, a slow diffusion process was used. The PDI compounds were dissolved in a good solvent, chloroform, then a bad solvent, methanol, was gently added on top of the chloroform solution, such that methanol slowly diffused into chloroform.9,24,18,31 This diffusion resulted in aggregation and/or crystallization of the PDI molecules into nanostructures with the size and shape dependent on the identity of the substituents. Optoelectronic Properties. The absorption spectrum of solvated PDI A, shown in Figure 3A (black spectrum), reveals the well-established absorption maxima at 526 nm, 489 nm, and 458 nm, attributed to the 0-0, 0-1, and 0-2 vibronic transitions, respectively.1,7, 9,14, 21,23,30,32 As mentioned above, the imide positions of PDI are nodes in the π orbital wave functions and therefore do not impact the electronic structure of the perylene core. Thus, all PDI molecules prepared have essentially the same absorption profile when fully solvated.10,11,14,16, 22,29,32

As expected, the absorption spectra of alkylated PDI molecules changes upon self-assembly. Side chain morphology can play a crucial role in π−π stacking distance, ultimately dictating intermolecular interactions and thus controlling the selfassembly process.14,22,23 For example, Würthner et al. found that substitution with linear alkyl chain (least steric demand) formed H-aggregates that are identified by a shift in absorption to ~550 nm.7 In compliment, when the imide positions are substituted with branched alkyl groups (higher steric demand), J-aggregates are formed, as observed by a shift in absorption to ~620 nm.7 The absorption spectrum of PDI A aggregates in suspension is similar to the solvated form with three vibronic transitions, but with a slight increase in absorbance at longer wavelengths (Figure 3A, red spectrum). Similar trends were observed in the absorption spectra of suspensions of aggregated PDI B, PDI 1’s, PDI 2’s, and PDI 3’s (Figure S18). The absorption bands emerging at longer wavelengths indicate strong π-π interactions and co-facial molecular stacking.7,11,14,33,34 The UV-Vis spectra of aggregated PDI molecules drop cast onto a glass slide were distinctly different from the

suspended aggregates, suggesting that solvated PDI molecules dominate the spectrum of suspended aggregates. The absorption profiles of many PDI derivatives in the solid state have been associated with an overall decrease in absorption intensity, broadening of vibronic bands, and formation of new bands at longer and shorter wavelengths, indicating the formation of both H- and J-aggregates.4,14,35 Likewise, the absorption spectrum of PDI A aggregates in the solid state (Figure 3A, purple spectrum) indicates formation of both H- and Jaggregates (absorption at 561 and 606 nm, respectively). Whereas branched alkyl chains can increase the distance between aromatic cores of neighboring PDI molecules,14,30,36 linear alkyl chains should facilitate self-assembly.9,14,23 The solid state UV-vis spectrum of aggregates of PDI B, which bears one branched alkyl substituent and one linear alkyl substituent, shows absorbance over the entire region, 400-550 nm, indicating weaker coupling interactions than those of PDI A (Figure 4C, purple spectrum). As both PDI A and PDI B bear only alkyl substituents, the difference in absorption spectra, and thus assembly are attributed only to symmetric versus asymmetric functionalization, and not to other van der Waals interactions (e.g. H-bonding). Pasaogullari et al. similarly observed different absorption spectra for symmetric and unsymmetric functionalized PDI aggregates, and attributed this to different intermolecular interactions.35 The absorption spectra of PDI 1’s, PDI 2’s, and PDI 3’s are impacted not just by differences in intermolecular interactions brought about by asymmetric functionalization, but also by other interactions, such as H-bonding. The solid state absorption spectra of aggregates of PDI 1-OH and PDI 2-OH have broadened absorption and increased absorbance in the far red region (>700 nm), indicating stronger intermolecular coupling, and both H- and J- aggregation.30 Of note, absorption at the longest wavelengths can be attributed to hydrogen bonding.7,30,35,37 In comparison, the solid state absorption spectrum of aggregates of PDI 3-OH is less red shifted (>600 nm), and may indicate that the longer alkyl chain prevents strong Hbonding. All PDI 1-OSiR3, PDI 2-OSiR3, PDI 3-OSiR3 (Figure

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

4A- C), show absorbance from ~450-600 nm, with a tailing into the red.

vent effects.11 The emission spectra of all PDI 1’s, PDI 2’s, and PDI 3’s resemble that of the PDI A, both in the solvated form, and for the suspensions of the aggregates (Figure 4D-F). Emission spectra in the solid state of assembled PDI’s were

Figure 4. UV-vis spectra of aggregated PDI molecules drop cast on a microscope slide. A) PDI A, PDI 1’s; B) PDI 2’s; C) PDI B, PDI 3’s. Fluorescence spectra (λex = 450 nm) of suspended PDI crystals D) PDI A, PDI 1’s; E) PDI 2’s; F) PDI B, PDI 3’s. Refer to Figure 2B for color code key.

The spectral shifts observed in the solid state are clearly influenced by the spacing between molecules particularly determined by the size/bulk of the substituents. Although, PDI A contains a short, branched alkyl chain, strong coupling interactions were observed that were not present in the spectra of PDI B, PDI 1-OSiR3, PDI 2-OSiR3, PDI 3-OSiR3 and PDI3-OH. Moreover, length of the alkyl chain significantly impacts the electronic spectra and influences hydrogen bonding interactions. The emission spectra of solvated and aggregated PDI can also be used to understand intermolecular interactions; the black trace in Figure 3B shows the emission spectrum of solvated PDI A and reveals the expected three vibronic transitions: 0-0, 0-1, and 0-2 bands, a mirrored image of the absorption spectrum.4, 14, 15, 18, 23, 24 The purple trace in Figure 3B shows the emission spectrum of a suspension of aggregates of PDI A, showing similar transitions to the solvated form, but with a different relative intensity. Specifically, emission intensity at longer wavelengths is observed and can be attributed to Jaggregate formation. The emission maxima of the longer wavelength bands blue shifts ~2 nm and is likely due to sol-

featureless, with no observable signal. Thermal properties. To evaluate the impact of functionalization on the thermal properties of the asymmetrically functionalized PDI molecules, thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) were used. Figures 5A-D show the weight loss profiles of the asymmetrically functionalized PDI molecules, organized by the amino alcohol used (Figure 5B = PDI 1’s, Figure 5C = PDI 2’s, and Figure 5D = PDI 3’s). All silylated PDI molecules are more thermally stable than the hydroxyl containing molecules, showing slight differences before rapid weight loss above 400 °C due to loss of alkyl chains (see Figure S19 for details), followed by another degradation step above 600 °C for the perylene core.38,39 All hydroxyl containing PDI compounds show a distinct, though slight weight loss transition at ~200 °C, which may indicate reactivity of hydroxyl groups or changes in hydrogen bonding. The thermotropic behavior of liquid crystal molecules, such as PDI, can be investigated by DSC;1,3,14,16,21,23,24,35 Figure 5E shows the thermal profile of PDI A, revealing an exothermic peak at 65.2 °C corresponding to crystallization (upon cooling), and an endothermic peak at 73.2 °C upon heat-

ACS Paragon Plus Environment

Page 4 of 9

Page 5 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 5. TGA weight loss profiles of: A) PDI A and PDI B; B) PDI 1’s; C) PDI 2’s; and D) PDI 3’s; DSC thermograms of: E) PDI A and PDI B, F) PDI 1’s; and G) PDI 2’s; and H) PDI 3’s. Refer to Figure 2B for color code key. ing, corresponding to melting. These data suggest that PDI A is crystalline. Figure 5E also shows the DSC trace of PDI B, and reveals only weak signals for crystallization and melting, as reported for some PDI derivatives with weakened or distorted molecular stacking, and is in agreement with the electronic spectra discussed above.14,35 Figures 5F-H show the DSC thermal profiles of PDI 1’s, PDI 2’s, and PDI 3’s, and all are nearly featureless. This does not support that the PDI molecules are crystalline. However, PDI 2-OTBDMS does show two distinct, though weak, transitions at 20.0 °C and -6.5°C indicating a crystalline nature (Figure 5G, pink), even though the absorption spectrum is similar to the other asymmetrically functionalized PDI molecules. For each sample, the third heating cycle is shown, so as to neglect any impact of thermal history, and heating was performed at 10 °C min-1 under nitrogen, as typical for PDI derivatives.1,6,16,18 Microscopy. A small drop of the suspended PDI aggregates was placed on copper grids covered with a thin carbon film for characterization by scanning electron microscopy (SEM) and transmission electron microscopy (TEM). Figure 6 shows the SEM images of the self-assembled PDI molecules (other SEM images are included in Figures S21-S34). PDI A and PDI B are symmetrically and asymmetrically functionalized with only alkyl substituents; from the SEM images, the assemblies are homogeneous and thread-like with smooth surfaces. They are microns long, with the diameter dictated by the alkyl substituents. In contrast, for the PDI 1’s, PDI 2’s, and PDI 3’s, both the identity of the amino alcohol and the identity of the trialkyl silyl group impact the shape of the structures formed. Within the family of PDI 1’s, crystals were relatively heterogeneous with tapered ends, except for PDI 1-OH which had relatively homogeneous surfaces. PDI 1-OTMS shows signs of nucleated assembly, indicating that aggregates of PDI may serve as nucleation sites for subsequent assembly of molecules, not observed for other compounds. In an attempt to pre-

vent the nucleated growth, the chloroform solution was filtrated multiple times before addition of methanol, but no difference was observed. As seen in Figure 6, structures of assembled PDI 1-OTBDMS and PDI 1-OTIPS were splintered, which may be attributed to defects within the self-assembled structure due to the proximity of the bulky trialkyl silyl group to the PDI core. PDI 2-OH assembled into blocky and short crystals that were < 30 µm long, yet PDI 2-OTMS, PDI 2OTBDMS, and PDI 2-OTIPS all assembled into structures with substantially higher aspect ratio. All PDI 3’s assembled into long structures with relatively homogeneous surfaces (in comparison to PDI 1’s and PDI 2’s), and were more similar to the dialkyl PDI derivatives. PDI 3-OH assembled into smooth structures 100’s of microns long, whereas PDI 3-OTMS showed structures of similar length, but with tapered ends. PDI 3-OTBDMS and PDI 3-OTIPS both show more ribbonlike structures that undulate, and in the case of PDI 3-OTIPS the ribbons are twisted (though not chiral). These results indicate that both hydroxyl and silyl ether substituted asymmetrically functionalized PDI molecules assemble into nanostructures using a solvent diffusion method. While all hydroxyl containing compounds can interact by hydrogen bonding, π−π stacking, and alkyl-alkyl interactions, their assembly is relatively well controlled and high aspect ratio structures are obtained. Overall, a longer alkyl linker between the silyl ether and PDI core allows for higher aspect ratio structures to be formed, likely due to better accommodation of the sterically demanding silyl groups. Although trialkyl silyl groups impact self-assembly, there is no universal trend (i.e., larger silyl groups do not necessarily lead to smaller structures); indeed, the shape and size of the assemblies formed are defined by both the trialkyl silyl group and the alkyl linker to the PDI core.

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 6. SEM images of symmetrically and asymmetrically functionalized PDI molecules after self-assembly using the solvent diffusion method. Expanded views of the samples and histograms are available in the supporting information (Figures S21-S34). Refer to Figure 2B for color code key. Crystallinity. TEM with electron diffraction has been used to determine the crystallinity of self-assembled small molecules.9,21,23,40 As shown in Figure S20, data quickly collected for PDI 1-OH, PDI 2-OH, and PDI 3-OTIPS show patterns consistent with highly crystalline structures, but are not sufficient for further analysis; moreover, these diffraction spots rapidly degraded into a haze ring pattern during analysis. Repeated attempts to collect meaningful crystallinity data by TEM diffraction were met with little success due to significant damage of the sample from the electron beam.9 All attempts to decrease voltage and adjust exposure time were not sufficient. In contrast, characterization of the PDI molecules by powder XRD revealed the samples were crystalline and/or polycrystalline (Figure S35). Conclusion Herein we have reported the synthesis and characterization of twelve different asymmetrically functionalized perylene diimide (PDI structures). These small molecules were synthesized, purified, and the chemical composition characterized by 1 H and 13C NMR, FTIR, and mass spectrometry. Using the solvent-diffusion method, these molecules were selfassembled into nanostructures. UV-Vis absorption spectra reveal a red shift in absorption after self-assembly, characteris-

tic of p-p interactions and cofacial stacking of the perylene core; emission spectra of suspended aggregates resembles that of solvated PDI, indicating that solvated molecules dominate the spectra. In contrast, the emission profile of the aggregates in the solid state are featureless. Characterization of the thermal properties of the molecules reveals that the silylated PDI’s are more stable than the PDI molecules bearing alcohol functionalities, and that all samples undergo rapid weight loss above 400 °C. The DSC thermograms of the asymmetrically functionalized PDI molecules are all featureless, except for PDI-2OTBDMS. However, powder XRD shows the samples are crystalline. Images of the nanostructures collected by SEM show that the size and shape of the assemblies are impacted by the length of the alkyl chain, as well as the identity of the end group functionality (alcohol or silyl ether). Thus, while asymmetric functionalization of PDI can be used to dictate the structure formed upon self-assembly, no overarching trend is identified for the set of molecules provided herein. Ongoing work focuses on identifying the location of the hydroxyl and silylated groups within the assemblies, as well as exploring other moieties to control the size, shape, and functionality of the structures.

ASSOCIATED CONTENT

ACS Paragon Plus Environment

Page 6 of 9

Page 7 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials Supporting Information. Experimental detail, supporting electron microscopy images, absorption and emission spectra. This material is available free of charge via the Internet at http://pubs.acs.org.

(13)

AUTHOR INFORMATION (14)

Corresponding Author * [email protected]

(15)

Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

(16)

Funding Sources The authors thank Case Western Reserve University College of Arts and Sciences for financial support. MALDI-TOF is supported by the NSF MRI-0821515. SEM and TEM was performed at the Swagelok Center for Surface Analysis of Materials (SCSAM) at CWRU. The authors acknowledge NSF MRI-1334048 for NMR instrumentation. K.M.H. acknowledges Texas Tech University for financial support.

References (1)

(2)

(3)

(4)

(5)

(6)

(7)

(8)

(9)

(10) (11)

(12)

(17)

(18)

(19)

Struijk, C. W.; Sieval, A. B.; Dakhorst, J. E. J.; Dijk, M. Van; Kimkes, P.; Koehorst, R. B. M.; Donker, H.; Schaafsma, T. J.; Picken, S. J.; Craats, A. M. Van De; et al. Liquid Crystalline Perylene Diimides : Architecture and Charge Carrier Mobilities. J. Am. Chem. Soc 2000, 122, 11057–11066. Palermo, V.; Liscio, A.; Gentilini, D.; Nolde, F.; Müllen, K.; Samori, P. Scanning Probe Microscopy Investigation of SelfOrganized Perylenetetracarboxdiimide Nanostructures at Surfaces: Structural and Electronic Properties. Small 2007, 3, 161–167. Fan, Y.; Ziabrev, K.; Zhang, S.; Lin, B.; Barlow, S.; Marder, S. R. Comparison of the Optical and Electrochemical Properties of Bi(perylene Diimide)s Linked through Ortho and Bay Positions. ACS Omega 2017, 2, 377–385. Amiralaei, S.; Uzun, D.; Icil, H. Chiral Substituent Containing Perylene Monoanhydride Monoimide and Its Highly Soluble Symmetrical Diimide: Synthesis, Photophysics and Electrochemistry from Dilute Solution to Solid State. Photochem. Photobiol. Sci. 2008, 7, 936–947. Cochrane, K. A.; Schiffrin, A.; Roussy, T. S.; Capsoni, M.; Burke, S. A. Pronounced Polarization-Induced Energy Level Shifts at Boundaries of Organic Semiconductor Nanostructures. Nat. Commun. 2015, 6, 1–8. Wu, N.; Zhang, Y.; Wang, C.; Slattum, P. M.; Yang, X.; Zang, L. Thermoactivated Electrical Conductivity in Perylene Diimide. J. Phys. Chem. Lett. 2017, 8, 292–298. Ghosh, S.; Li, X. Q.; Stepanenko, V.; Würthner, F. Control of H- and J-Type π Stacking by Peripheral Alkyl Chains and SelfSorting Phenomena in Perylene Bisimide Homo- and Heteroaggregates. Chem. - A Eur. J. 2008, 14, 11343–11357. 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, 2386– 2407. Briseno, A. L.; Mannsfeld, S. C. B.; Reese, C.; Hancock, J. M.; Xiong, Y.; Jenekhe, S. a; Bao, Z.; Xia, Y. Perylenediimide Nanowires and Their Use in Fabricating Field-Effect Transistors and Complementary Inverters. Nano Lett. 2007, 7, 2847–2853. Kozma, E.; Catellani, M. Perylene Diimides Based Materials for Organic Solar Cells. Dye. Pigment. 2013, 98, 160–179. Chen, Y.; Feng, Y.; Gao, J.; Bouvet, M. Self-Assembled Aggregates of Amphiphilic Perylene Diimide-Based Semiconductor Molecules: Effect of Morphology on Conductivity. J. Colloid Interface Sci. 2012, 368, 387–394. Lee, S. K.; Zu, Y.; Herrmann, A.; Geerts, Y.; Mullen, K.; Bard,

(20)

(21)

(22)

(23)

(24)

(25)

(26)

(27)

(28)

(29)

A. J. Electrochemistry, Spectroscopy and Electrogenerated Chemiluminescence of Perylene, Terrylene, and Quaterrylene Diimides in Aprotic Solution. J. Am. Chem. Soc. 1999, 121, 3513–3520. Zhou, Z.; Brusso, J. L.; Holdcroft, S. Directed Growth of 1D Assemblies of Perylene Diimide from a Conjugated Polymer. Chem. Mater. 2010, 22, 2287–2296. Balakrishnan, K.; Datar, A.; Naddo, T.; Huang, J.; Oitker, R.; Yen, M.; Zhao, J.; Zang, L. Effect of Side-Chain Substituents on Self-Assembly of Perylene Diimide Molecules: Morphology Control. J. Am. Chem. Soc. 2006, 128, 7390–7398. Wicklein, A.; Kohn, P.; Ghazaryan, L.; Thurn-Albrecht, T.; Thelakkat, M. Synthesis and Structure Elucidation of Discotic Liquid Crystalline Perylene Imide Benzimidazole. Chem. Commun. 2010, 46, 2328–2330. Wicklein, A.; Lang, A.; Muth, M.; Thelakkat, M. Swallow-Tail Substituted Liquid Crystalline Perylene Bisimides: Synthesis and Thermotropic Properties. J. Am. Chem. Soc. 2009, 131, 14442–14453. Hu, J.; Kuang, W.; Deng, K.; Zou, W.; Huang, Y.; Wei, Z.; Faul, C. F. J. Self-Assembled Sugar-Substituted Perylene Diimide Nanostructures with Homochirality and High Gas Sensitivity. Adv. Funct. Mater. 2012, 22, 4149–4158. Ren, X.; Sun, B.; Tsai, C.-C.; Tu, Y.; Leng, S.; Li, K.; Kang, Z.; Horn, R. M. Van; Li, X.; Zhu, M.; et al. Synthesis, SelfAssembly, and Crystal Structure of a Shape-Persistent Polyhedral-Oligosilsesquioxane-Nanoparticle-Tethered Perylene Diimide. J. Phys. Chem. B 2010, 114, 4802–4810. Rajaram, S.; Armstrong, P. B.; Kim, B. J.; Fre´chet, J. M. J. Effect of Addition of a Diblock Copolymer on Blend Morphology and Performance of Poly(3hexylthiophene):Perylene Diimide Solar Cells. Chem. Mater. 2009, 21, 2008–2010. Zhang, Y.; Zheng, Y.; Xiong, W.; Peng, C.; Zhang, Y.; Duan, R.; Che, Y.; Zhao, J. Morphological Transformation between Nanocoils and Nanoribbons via Defragmentation Structural Rearrangement or Fragmentation-Recombination Mechanism. Sci. Rep. 2016, 6 1–8. 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, 10496–10497. Sun, J. P.; Hendsbee, A. D.; Dobson, A. J.; Welch, G. C.; Hill, I. G. Perylene Diimide Based All Small-Molecule Organic Solar Cells: Impact of Branched-Alkyl Side Chains on Solubility, Photophysics, Self-Assembly, and Photovoltaic Parameters. Org. Electron. physics, Mater. Appl. 2016, 35 151–157. Boobalan, G.; Imran, P. K. M.; Nagarajan, S. Luminescent OneDimensional Nanostructures of Perylene Bisimides. Spectrochim. Acta - Part A Mol. Biomol. Spectrosc. 2013, 113, 340– 345. Islam, M. R.; Sundararajan, P. R. Self-Assembly of a Set of Hydrophilic–solvophobic–hydrophobic Coil–rod–coil Molecules Based on Perylene Diimide. Phys. Chem. Chem. Phys. 2013, 15, 21058–21069. Zhang, X.; Chen, Z.; Wu, F. Morphology Control of Fluorescent Nanoaggregates by Co-Self-Assembly of Wedge- and Dumbbell-Shaped Amphiphilic Perylene Bisimides. J. Am. Chem. Soc. 2007, 129, 4886–4887. Mondal, S.; Lin, W. H.; Chen, Y. C.; Huang, S. H.; Yang, R.; Chen, B. H.; Yang, T. F.; Mao, S. W.; Kuo, M. Y. SolutionProcessed Single-Crystal Perylene Diimide Transistors with High Electron Mobility. Org. Electron. 2015, 23, 64–69. Che, Y.; Datar, A.; Balakrishnan, K.; Zang, L. Ultralong Nanobelts Self-Assembled from an Asymmetric Perylene Tetracarboxylic Diimide. J. Am. Chem. Soc. 2007, 129, 7234– 7235. Ma, X.; Zhang, Y.; Zheng, Y.; Tao, X.; Che, Y.; Zhao, J. Highly Fluorescent One-Handed Nanotubes Assembled from a Chiral Asymmetric Perylene Diimide. Chem. Commun. 2015, 51, 4231–4233. Liu, X.; Zhang, Y.; Pang, X.; Yue, E.; Zhang, Y.; Yang, D.; Tang, J.; Li, J.; Che, Y.; Zhao, J. Nanocoiled Assembly of Asymmetric Perylene Diimides: Formulation of Structural Factors. J. Phys. Chem. C 2015, 119, 6446–6452.

ACS Paragon Plus Environment

Chemistry of Materials (30)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(31)

(32)

(33)

(34)

(35)

Weintraub, M. T.; Xhakaj, E.; Austin, A.; Szarko, J. M. The Effects of Donor : Acceptor Intermolecular Mixing and Acceptor Crystallization on the Composition Ratio of Blended, Spin Coated Organic Thin Films. J. Mater. Chem. C 2016, 4, 7756–7765. Bu, L.; Pentzer, E.; Bokel, F. A.; Emrick, T.; Hayward, R. C. Growth of Polythiophene / Perylene Tetracarboxydiimide Donor/Acceptor Shish-Kebab Nanostructures by Coupled Crystal Modi Fi Cation. ACS Nano 2012, 6, 10924–10929. Kennehan, E. R.; Grieco, C.; Brigeman, A. N.; Doucette, G. S.; Rimshaw, A.; Bisgaier, K.; Giebink, N. C.; Asbury, J. B. Using Molecular Vibrations to Probe Exciton Delocalization in Films of Perylene Diimides with Ultrafast Mid-IR Spectroscopy. Phys. Chem. Chem. Phys. 2017, 19, 24829–24839. LING ZANG, Y. C.; MOORE, J. S. One-Dimensional SelfAssembly of Planar π- Conjugated Molecules: Adaptable Building Blocks for Organic Nanodevices. Acc. Chem. Res. 2008, 41, 1596–1608. Sarbu, A.; Biniek, L.; Guenet, J.-M.; Mésini, P. J.; Brinkmann, M. Reversible J- to H-Aggregate Transformation in Thin Films of a Perylenebisimide Organogelator. J. Mater. Chem. C 2015, 3, 1235–1242. Pasaogullari, N.; Icil, H.; Demuth, M. Symmetrical and

(36)

(37)

(38)

(39)

(40)

Unsymmetrical Perylene Diimides: Their Synthesis, Photophysical and Electrochemical Properties. Dye. Pigment. 2006, 69, 118–127. 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, 3715–3725. Würthner, F.; Thalacker, C.; Sautter, a; Schärtl, W.; Ibach, W.; Hollricher, O. Hierarchical Self-Organization of Perylene Bisimide--Melamine Assemblies to Fluorescent Mesoscopic Superstructures. Chemistry 2000, 6, 3871–3886. Türkmen, G.; Erten-Ela, S.; Icli, S. Highly Soluble Perylene Dyes: Synthesis, Photophysical and Electrochemical Characterizations. Dye. Pigment. 2009, 83, 297–303. Nagao, Y.; Misono, T. Synthesis and Properties of N-Alkyl-N′Aryl-3,4:9,10-Perylenebis(dicarboximide). Dye. Pigment. 1984, 5, 171–188. Zhao, L.; Yang, W.; Ma, Y.; Yao, J.; Li, Y.; Liu, H. Perylene Nanotubes Fabricated by the Template Method. Chem. Commun. 2003, 19, 2442–2443.

ACS Paragon Plus Environment

Page 8 of 9

Page 9 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Table of Contents artwork

ACS Paragon Plus Environment

9