UNIVERSIDAD COMPLUTENSE DE MADRID FACULTAD DE CIENCIAS QUÍMICAS TESIS DOCTORAL Chiral Molecular Nanographenes: Synthesis and Properties Nanografenos Moleculares Quirales: Síntesis y Propiedades MEMORIA PARA OPTAR AL GRADO DE DOCTORA PRESENTADA POR Patricia Izquierdo García DIRECTORES Prof. Nazario Martín León Dr. Jesús Manuel Fernández García © Patricia Izquierdo García, 2024 UNIVERSIDAD COMPLUTENSE DE MADRID FACULTAD DE CIENCIAS QUÍMICAS TESIS DOCTORAL Chiral Molecular Nanographenes: Synthesis and Properties Nanografenos Moleculares Quirales: Síntesis y Propiedades MEMORIA PARA OPTAR AL GRADO DE DOCTORA EN QUÍMICA ORGÁNICA presentada por Patricia Izquierdo García DIRECTORES Prof. Nazario Martín León Dr. Jesús Manuel Fernández García AGRADECIMIENTOS Primero, agradecer a la Universidad Complutense de Madrid, a la Facultad de Química y al Departamento de Química Orgánica por la formación que he podido adquirir a lo largo de mi carrera en la química, desde los estudios de grado y máster, hasta esta última etapa de formación de la tesis doctoral que, además, he podido realizar con un Contrato Predoctoral de Personal Investigador en Formación UCM- Harvard. Nazario y Chus, gracias por convertir estos años en un libro de aventuras, concretamente en uno de esos que al terminar te dejan una especie de resaca emocional. Pensándolo bien, el título de esta tesis tal vez debería ser algo como “Unexpected and Exciting Adventures With Molecular Nanographenes”, pero esa discusión ya la dejamos para otro día. Es difícil resumir en unas pocas palabras la admiración y respeto que siento por los dos. Gracias por acompañarme y darme las herramientas para ser capaz de enfrentarme a los desafíos que la química tenía preparados. Gracias por impulsarme cuando las fuerzas flaqueaban y por confiar en mí. Gracias por ser tan humanos y cercanos, por encontrar siempre las palabras amables y también las más sinceras. Gracias por las infinitas horas de fructíferas reuniones discutiendo de química y quiralidad, con las que pudimos mantener la cordura cuando parecía que el mundo iba a la deriva. Espero seguir desafiando a la mente con la ciencia tal y como me habéis enseñado, siempre buscando nuevos objetivos, sin perder la curiosidad, la creatividad y el entusiasmo. Gracias por todo, sois referentes para mí, como científicos, como mentores y como personas. Así, y como no podría ser de otra manera, la Patri que termina esta etapa ha aprendido muchas lecciones que la acompañarán siempre… Me gustaría agradecer al Dr. Jarad Mason de la Universidad de Harvard el brindarme la oportunidad de trabajar en su laboratorio. Durante los tres meses de estancia en el grupo de investigación tuve la posibilidad de asistir a conferencias del más alto nivel y de conocer a personas totalmente entusiasmadas con la ciencia y siempre dispuestas a ayudar, mencionando especialmente a Joy, Faith, Quichen, Vidhya, Yukyung, Dan, Rahil y Miranda. A todas las personas que han permitido mejorar los trabajos que se describen en esta tesis doctoral. A los colaboradores Juan Casado, Samara Medina, Irena Stará, Jeanne Crassous, Sergio Ramírez y David Fresnadillo, por sus aportaciones en el estudio de los nanografenos moleculares bicapa. Al colaborador Israel Fernández por realizar los cálculos teóricos que soportan los resultados de la formación de espironanografenos y nanografenos helicoidalmente organizados. Por último, gracias especialmente a Josefina Perles, por arrojar luz con la difracción de rayos-X en todos los trabajos desarrollados. Por acompañarnos concienzudamente en todas y cada una de las historias que hemos contado, incluso en aquellas que más bien han parecido películas de Hitchcock, en las que la esperanza podría llegar a perderse, pero siempre lo intentábamos una vez más. Muchas gracias a Lola, Ángel y Elena, que desde el CAI de RMN nos hacen el trabajo más sencillo. Gracias a Lola, especialmente por hacer de atleta esperando a recibir el testigo del tubo de resonancia con el que cada minuto contaba, porque nuestro querido nanografeno isomerizaba. Y, por supuesto, gracias a Elena, por sacar adelante con determinación y paciencia todos los desafíos que a priori pudieran parecer locas ideas mías. Sergio, mi querido Sergio, ¿qué habría sido de mí sin ti estos años? Hay preguntas que es mejor no hacerse porque, por suerte, te he tenido a mi lado. Gracias una vez más, por conocerme, por entenderme, por hacerme reír y apoyarme cuando tocaba llorar o rabiar. Eres la definición de amigo, pero todo esto ya lo sabes. Como esto es mi libro y puedo poner “lo que quiera” (y por evitar aquello de que las palabras se las lleva el viento), quiero aprovechar para decirte que vales muchísimo, que conseguirás todo lo que te propongas, como has hecho hasta ahora. Solo espero que seas consciente de ello y que creas en ti y, si no lo consigues, me llamas y yo te lo recuerdo, porque lo tengo claro. Y como dice La Pelae para darnos fuerza, “¡¡métele castaña!!”, ya sabes cómo sigue… Manu, ¡la de cosas que han pasado en lo que han parecido dos días!! Empezamos justamente en el cambio generacional del laboratorio, nos dejaron siendo dos novatillos y, sin querer, nos convertimos en algo parecido a “papá” y “mamá”. Ahora nos vamos nosotros, pero creo que los niños ya son mayores y estarán bien. Gracias por estos años en el laboratorio, por estar dispuesto a hablar de química, pero también a hacer el mono si era lo que tocaba ese día. Sé que te irá bien en lo que está por venir, lo único, no pongas música del infierno o de tienda de ropa el primer día por si hay gente de tímpano sensible, prueba primero con algo como Florence... Mis tres españolas en Boston, Carlota, María y Natalia, gracias por los paseos en bici, los atardeceres en la explanada, los días en la playa, las noches de juegos de mesa (un poco hipercompetitivos) y las tardes de peli/siesta. Pero, sobre todo, gracias por hacerme sentir en casa a 5000 km de la mía. Con vosotras volví a ser yo misma cuando estaba un poco perdida en mi cabeza, y os lo agradeceré siempre. Por supuesto, a la gente del grupo, las personas con las que se han forjado recuerdos inolvidables, como cantar a pleno pulmón desde Bisbal a los Rebujitos los viernes por la tarde cuando el agotamiento llevaba a todos a delirar un poco, excepto a Manu, que estaría haciendo alguna columna. Gracias por estos años al inesperablemente sensible Sergio, a Gema, que coge las riendas cuando hace falta, a Jaime, Juan, Jenni, Arturo, Alex, Carlos, Antonio, Carlos, Valeria, Jesús, Marina, Sergio, Diego, Jesús, Justo, Javi, Agus… Gracias a Inés, que desde el principio me contó con total sinceridad lo que podría suponer hacer una tesis doctoral y que, por suerte, ha estado para apoyarme. A Paul por enseñarme a dar mis primeros pasos en el laboratorio. A los profesores del grupo, José, Laura, Salvatore, Beti y, especialmente, a Mª Ángeles por el continuo apoyo a lo largo de estos años. Gracias también a Ana, la persona que hace que el complejo engranaje del grupo funcione sin problemas echando una mano siempre que es necesario. A la gente del departamento, Jon, Vero, Anabel, Román, Paola, Cristina, Lucía, Adrián, María, Pablo, Mario, Matías… gracias por las tardes de desconectar y pasar un buen rato. Gracias, Andrea, por acogerme desde el primerísimo día, eres de esas personas a las que se les nota mucho lo de ser buena gente. Ángeles, has sabido entenderme como casi nadie, conocerme más de lo que yo pensaba, gracias por las charlas motivacionales y los momentos de desconexión. Tamara, solo me queda decirte que lo mejor está por llegar y, ya sabes, llama, grita si me necesitas. Marcos, mi hermano encubierto, eres todo vida, con tus ideas locas y tu capacidad para la investigación no habrá quién te pare. Por último… Queridos papá y mamá, lo que veis en este libro es lo que he hecho los últimos años jugando a los cacharritos y dibujando. Porque la niña salió de ciencias en la casa de las artes y las letras… En este tiempo he intentado hacer todo tal y como me habéis enseñado siempre, sin rendirme, con creatividad, paciencia, esfuerzo, determinación, valentía y, sobre todo tratando de disfrutar cada día. También he intentado ser fuerte como la abuela Trini. He seguido los consejos del abuelo Antonio, de estudiar todo lo que pudiera, porque el saber no ocupa lugar. Y, por supuesto, he trabajado duro y lo seguiré haciendo para ser una persona de provecho, como decía el abuelo Nicolás. Lo de ser cabezota me salía sin querer. Begoña, mi loquita hermana pequeña, mi QueenB, gracias por entenderme y estar cuando las cosas se torcían. Considerando tus esfuerzos por saber lo que es un anillo aromático, te presento el libro ideal, porque trata única y exclusivamente de muchos aromáticos. Lara, gracias por hacerme tan feliz, por entenderme, por estar siempre en mi equipo, por ser siempre mi apoyo incondicional. Gracias por ser mi familia, por hacerme quien soy, sin dejarme olvidar de dónde vengo y apoyarme siempre, aunque en el fondo no tenga del todo claro a dónde voy… PUBLICATIONS o P. Izquierdo-García, J. M. Fernández-García, I. Fernández, J. Perles, N. Martín, Helically Arranged Chiral Molecular Nanographenes, J. Am. Chem. Soc. 2021, 143, 11864. o P. Izquierdo-García, J. M. Fernández-García, J. Perles, I. Fernández, N. Martín, Electronic Control of the Scholl Reaction: Selective Synthesis of Spiro vs Helical Nanographenes, Angew. Chem. Int. Ed. 2023, 62, e202215655. o P. Izquierdo-García, J. M. Fernández-García, S. Medina Rivero, M. Šámal, J. Rybáček, L. Bednárová, S. Ramírez-Barroso, F. J. Ramírez, R. Rodríguez, J. Perles, D. García-Fresnadillo, J. Crassous, J. Casado, I. G. Stará, N. Martín, Helical Bilayer Nanographenes: Impact of the Helicene Length on the Structural, Electrochemical, Photophysical, and Chiroptical Properties, J. Am. Chem. Soc. 2023, 145, 11599. o J. M. Fernández-García, P. Izquierdo-García, M. Buendía, S. Filippone, N. Martín, Synthetic chiral molecular nanographenes: the key figure of the racemization barrier, Chem. Commun. 2022, 58, 2634. Table of Contents SUMMARY................................................................................................................v RESUMEN............................................................................................................. xiii INTRODUCTION.....................................................................................................1 The first 2D atomic crystal: a honeycomb patterned carbon layer..........................1 Trends and Recent Developments ...........................................................................4 Bilayer graphene: twist to superconductivity ......................................................4 Graphene molecules: Nanographenes .................................................................6 1 Chapter 1.......................................................................................................... 11 1.1 Introduction to the Synthesis of Nanographenes ................................ 11 1.1.1 Key Reactions for the π-Extension...................................................... 11 1.1.2 Graphitization ......................................................................................18 1.1.3 Scholl Cyclodehydrogenation .............................................................19 1.2 Objectives ...............................................................................................31 1.3 Results and Discussion...........................................................................33 1.3.1 Unexpected Synthesis of Spironanographenes ....................................33 1.3.2 Characterization of Spironanographene 10 .........................................37 1.3.3 Mechanism of the Scholl reaction .......................................................47 1.3.4 Electronic Effects Controlling the Scholl Reaction.............................48 1.3.5 Electrochemical and Photophysical Properties....................................50 1.4 Conclusions.............................................................................................57 2 Chapter 2..........................................................................................................61 2.1 Introduction to Chirality in Nanographenes .......................................61 2.1.1 From Two-Dimensionality to Chirality ...............................................63 2.1.2 Chiral Nanographenes and Isomerization Barriers..............................64 2.1.3 Alternative Approaches for Chirality in PAHs ....................................68 2.2 Objectives ...............................................................................................71 2.3 Results and Discussion...........................................................................73 2.3.1 Synthesis of Helically Arranged Nanographenes ................................73 2.3.2 Unexpected Chirality...........................................................................75 2.3.3 Molecular Dynamics: Asymmetric Derivatives 17b,c.........................77 2.3.4 No C−H ··· π Interactions, No Helicity, No Chirality .........................81 2.3.5 Electrochemical and Photophysical Properties ................................... 83 2.3.6 Photophysical Properties of Helically Arranged Nanographenes ....... 86 2.4 Conclusions ............................................................................................. 87 3 Chapter 3. ........................................................................................................ 91 3.1 Introduction to Bilayer Nanographenes.............................................. 91 3.1.1 Van der Waals Bilayer Molecular Nanographenes.............................. 91 3.1.2 Bilayers from Fused Radicals ............................................................. 94 3.1.3 Covalently Linked Bilayers ................................................................ 95 3.2 Objectives............................................................................................. 101 3.3 Results and Discussion........................................................................ 103 3.3.1 Synthesis of Helical Bilayer Nanographenes.................................... 103 3.3.2 Electrochemical and Spectroelectrochemical Properties .................. 105 3.3.3 Photophysical Properties................................................................... 107 3.3.4 Chiroptical Properties of HBNGs ......................................................111 3.4 Conclusions ...........................................................................................115 4 Chapter 4 ........................................................................................................119 4.1 Introduction to Stereochemical Control in Nanographenes.............119 4.1.1 Enantioselective Synthesis of Nanographenes...................................119 4.1.2 Chirality Transference: from Atropisomerism to Helical Chirality .. 120 4.2 Objectives............................................................................................. 125 4.3 Results and Discussion: oxa-Helical Bilayer Nanographenes ......... 127 4.3.1 Enantiospecific Synthesis of oxa-Helical Bilayer Nanographenes... 127 4.3.2 Electrochemical properties ............................................................... 133 4.3.3 Photophysical Properties................................................................... 135 4.3.4 Chiroptical properties ....................................................................... 136 4.4 Conclusions .......................................................................................... 137 EXPERIMENTAL SECTION ................................................................................ 141 APPENDIX 1. Chapter 1........................................................................................ 165 APPENDIX 2. Chapter 2........................................................................................ 175 APPENDIX 3. Chapter 3........................................................................................ 181 APPENDIX 4. Chapter 4........................................................................................ 193 APPENDIX 5. Spectra ........................................................................................... 201 REFERENCES ....................................................................................................... 257 Abbreviations i ABBREVIATIONS 1D: 1H-NMR 13C-NMR 19F-NMR 2D Abs AFM BCPL BINOL Cat CD COSY CPL Cu-TMDA DBPP DBU DCE DCM DDQ DEPT DFT DIPA DMF E0-0 E1 1/2ox/red e.e. Emi Fc/Fc+ FT-IR gabs/lum HBC HBNG HMBC HMQC HOMO HPLC HRMS INT IUPAC LUMO M m, MTDM MALDI-ToF Monodimensional Proton Nuclear Magnetic Resonance Carbon Nuclear Magnetic Resonance Fluorine Nuclear Magnetic Resonance Bidimensional Absorption Atomic Force Microscope Circularly Polarized Luminiscence Brightness 1,1′-Bi-2-naphthol Catalyst Circular Dichroism COrrelated SpectroscopY Circularly Polarized Light Di-μ-hydroxo-bis[(N,N,N′,N′-tetramethylethylenediamine)copper(II)] chloride Dibenzo[fg,ij]phenanthro-[9,10,1,2,3-pqrst]pentaphene 1,8-Diazabicyclo[5.4.0]undec-7-ene Dichloroethane Dichloromethane 2,3-Dichloro-5,6-dicyano-1,4-benzoquinone Distortionless Enhancement by Polarization Transfer Density Functional Theory Diisopropylamine N,N-Dimetilformamide Optical band gap First half wave oxidation/reduction potential Enantiomeric excess Emission Ferrocene/Ferrocenium Fourier Transform Infrared Spectroscopy Absorption and luminescence dissymmetry factor Hexa-peri-hexabenzocoronene Helical Bilayer Nanographene Heteronuclear Multiple Bond Correlation Heteronuclear Multiple Quantum Correlation Highest Occupied Molecular Orbital High Performance Liquid Chromatography High Resolution Mass Spectrometry Intermediate International Union of Pure and Applied Chemistry Lowest Unoccupied Molecular Orbital Molar Magnetic Dipole Transition Moment Matrix-Assisted Laser Desorption/Ionization-Time-of-Flight Abbreviations ii NBO NICS NIR NOESY [Ox] PAH r.t. SOPT tBu-HBC TD-DFT TFA TfOH THF Tol UV-vis V vs. ε θ λabs/emi μ, ETDM Φem Natural bond order Nuclear Independent Chemical Shift Near Infrared Nuclear Overhauser Effect Spectroscopy Oxidation Polyaromatic hydrocarbons Room temperature Second order perturbation theory Hexa-tert-butyl-hexa-peri-hexabenzocoronene Time-Dependent Density Functional Theory Trifluoroacetic acid Trifluoromethanesulfonic acid or triflic acid Tetrahydrofuran Toluene Ultraviolet-visible Volts versus Molar Absorption Coefficient Angle between ETDM and MTDM Absorption/Emission wavelength Electric Dipole Transition Moment Emission Quantum Yield Summary Summary v SUMMARY Chapter 1 The Scholl reaction is key in the synthesis of molecular nanographenes, as it leads to graphitization of the precursor polyarenes. The conditions for carrying out the reaction involve the use of Lewis acids or Brønsted acids in combination with oxidants such as DDQ (2,3-dichloro-5,6-dicyano-1,4-benzoquinone), and the mechanism involves the formation of arenium-cation or radical-cation intermediates. However, it is difficult to predict the operating mechanism and, sometimes, leads to unexpected compounds due to the formation of alternative intermediates or rearrangements. In this chapter, the unexpected synthesis of spironanographenes and the control over the Scholl reaction, by means of electronic effects on the structure, to obtain a helical nanographene, are described. The Scholl reaction on an anthracene-based polyarene results in the formation of spirocycles connecting the graphitized units to the central anthracene. As revealed by theoretical calculations, this occurs due to the formation of very stable trityl-cation intermediates, at positions 9 and 10 of the anthracene. Depending on the Scholl reaction conditions, it has been possible to control the number of graphitization closures, thus describing the first example in which the formation of spirocycles does not interrupt the graphitization. Pursuing the main objective of synthesizing a helically arranged nanographene consisting of two graphitized units (dibenzo[fg,ij]phenanthro-[9,10,1,2,3pqrst] pentaphene, DBPP) connected through a central anthracene, the Scholl reaction was controlled by the introduction of eight fluorine atoms. The electron-withdrawing character of the fluorine atoms destabilizes the formation of the trityl cation at the Summary vi positions 9 and 10 of the anthracene, causing the reaction to evolve through the arenium cation intermediates, thus, giving rise to the expected graphitized structure. In addition to the synthesis and structural characterization of the new molecular nanographenes, the characterization of the optoelectronic properties has been carried out. Cyclic voltammetry measurements of spironanographene showed higher absolute oxidation and reduction potentials (E1 1/2ox= 0.85 V, E1 1/2red= −2.56 V) than those corresponding to tBu-HBC (E1 1/2ox= 0.67 V and E1 1/2red= −2.27 V), which evidences the disruption of conjugation at the sp3 spiranic carbons connecting the graphitized units to the anthracene core. This structural feature has also an impact on the optical properties of the absorption and emission spectra, whose bands are blue shifted compared to those of tBu-HBC. The electronic properties of the helically arranged nanographene also revealed the lower π-extension of the graphitized units (DBPP), with respect to the tBu-HBC, showing a first oxidation potential at E1 1/2ox= 0.87 V. However, the acceptor character is stronger than that of tBu-HBC, as the first reduction potential (E1 1/2red= −1.69 V) is highly influenced by the localization of the LUMO on the anthracene electron- withdrawing core. This behavior is also reflected in the emission spectrum, in which a red shift of the bands is observed, compared to tBu-HBC. Interestingly, this structure appears to exhibit a dual emission in which the two fluorophores (DBPPs and octafluoroanthracene) seem to give rise to the corresponding bands independently. Chapter 2 Most strategies for inducing chirality in nanographenes involve deviation from planarity by introducing topological defects. Thus, non-hexagonal cycles and helical features are the most extensively studied topological defects. However, the possibility of inducing atropisomerism-like axial chirality, in which the chirality arises from the restriction of rotation of the structural units around a single bond, has been slightly studied. Therefore, in this chapter a new strategy to induce atropisomerism in nanographenes is described. In addition, the molecular dynamics, directly related to the stability of the chiroptical properties of helically arranged nanographenes, is studied. With the aim of simplifying the system to carry out the chirality study, nanographenes based on two graphitized units (DBPP) connected through a 1,4-disubstituted tetrafluorobenzene core have been synthesized, resembling the structure obtained in the previous chapter based on related octafluoroanthracene. Summary vii The initial approach consisted on the synthesis of a nanographene in which both, the substituents on the DBPP are the same, and the central core is substituted with four fluorine atoms. Subsequent modification of the substituents (in the DBPP and the central core) would lead to terphenyl-like atropisomerism. However, once the structure in which the DBPPs are substituted with five tert-butyl groups and the central benzene with four fluorine atoms in the 2,3,5,6 positions was obtained, two unexpected enantiomers were observed in the crystal structure. DFT calculations revealed the formation of C−H···π interactions between the DBPP units. These interactions occur as a result of the hindered coplanarity between the DBPPs and set the structure with a helical arrangement of the subunits, giving rise to two non- superimposable mirror images. The enantiomers were observed by chiral HPLC; however, they could not be isolated due to the racemization observed at room temperature. For the study of the isomerization barrier, two derivatives were prepared, in which one tert-butyl group on each DBPP was exchanged for a OMe group or a hydrogen atom. This modification involves the formation of two NMR-distinguishable nearly degenerate diastereomers. From a sample enriched in one of the diastereomers it was possible to monitor the evolution to equilibrium (1:1 ratio), which allowed us to calculate the value of the isomerization barrier, Δ𝐺⧧= 24.6 kcal·mol−1 at 40 ºC, and a half-time life of τ1/2 = 107 min. Furthermore, with these experiments it was possible to confirm that the 90° rotation of the tetrafluorobenzene unit was the reason for the isomerization process. Additionally, the electrochemical properties were studied by cyclic voltammetry and square wave voltammetry. The tetrafluorobenzene-based nanographenes are weaker electron-donor and electron-acceptors than tBu-HBC, which is in agreement with the lower π-extension of the DBPP units and the weak conjugation along the structure. The same trend was confirmed by absorption and emission spectra measurements, which showed blue-shifted bands with respect to tBu-HBC. Summary viii Chapter 3 In this chapter, the synthesis and characterization of two helical bilayer nanographenes (HBNGs) with different number of ortho-fused rings giving rise to two helicenes of different lengths, [9] and [11]helicene ([9]HBNG and [11]HBNG), is described. These structures, together with the one previously published by our research group [10]HBNG (ACIE 2018), form a family whose structural features determine the variation of properties. Commonly, the properties of molecular nanographenes can be modulated by varying the π-extension or by introducing defects in the hexagonal lattice (as described in Chapter 2). However, in this chapter we describe the impact of a new structural feature, the overlapping degree between two facing layers with intramolecular π,π- interaction. Therefore, this chapter details the systematic study of the optoelectronic and chiroptical properties of these helical bilayer nanographenes. The length of the helicene determines the overlapping degree between the graphitized layers fused at the ends, with [9]HBNG being the most overlapped (26 rings involved in π-π interactions) and [11]HBNG (10 rings involved in π-π interactions) being the least overlapped. Moreover, in these structures the π-extension varies with the number of rings in the helicene, with [9]HBNG having the lowest π-extension and [11]HBNG the highest π-extension. Taking these structural features into account, it was possible to study and understand the impact of the overlapping degree on the properties. The strongest donor character is exhibited by the lower π-extended nanographene [9]HBNG. This behavior is attributed to a mixed valence effect whereby the cation and the radical cation (oxidized species) are stabilized between the graphitized layers (greater overlapping degree). In addition, this structure [9]HBNG shows the most red-shifted emission bands, evidencing the communication of the layers through space by π-π interactions. Interestingly, the alignment of the electric and magnetic vectors makes [9]HBNG the nanographene with the best chiroptical properties, with an absorption and emission Higher overlapping Larger π-extension Stronger donor character [9]HBNG [10]HBNG [11]HBNG ACIE 2018 Summary ix factor of circularly polarized light (gabs and glum) of 3.6·10−2, one of the highest reported for carbon structures. Chapter 4 The growing interest in helical bilayer nanographenes, mainly due to their unique overlapping characteristics and remarkable chiroptical properties (Chapter 3), makes it necessary to search for new strategies to improve synthetic accessibility. In this chapter we describe an easily scalable synthetic strategy for chiral bilayer nanographenes that avoids the costly separation of enantiomers by chiral HPLC. The synthesis starts from the commercially available compound 7-bromo-2-naphthol, whose reactivity allows the production of 7,7'-disubstituted BINOL derivatives. The configurational stability of BINOL (with axial chirality), and the reactivity of the naphthol groups allowed the esterification with enantiomerically pure canforsulfonyl chloride for the chemical resolution of the enantiomers of the BINOL-based polyarene. Subsequently, the Scholl reaction was performed, in which cyclodehydrogenation (graphitization) and enantiospecific cyclodehydration of the naphthol groups occurs to give oxa[9]HBNG with chirality transference from axial to helical. In addition, the characterization of the optoelectronic and chiroptical properties has been carried out. By means of cyclic voltammetry, the structure oxa[9]HBNG (22 rings involved in the intramolecular π-π interactions) presents a slightly weaker donor character, E1 1/2ox= 0.37 V, than the bilayer [9]HBNG (26 rings involved in the intramolecular π-π interactions), E1 1/2ox= 0.35 V, evidencing the influence of the Summary x overlapping degree (lower overlapping, weaker donor character). Moreover, the most prominent redox variation is observed in the acceptor character, the presence of the oxygen atom in the helicene causes the LUMO to localize in the graphitized layers, increasing the reduction potential, E1 1/2red= −2.27 V, to values similar to those of tBu- HBC, E1 1/2red= −2.24 V. In addition, the emission spectrum of oxa[9]HBNG shows well-structured blue-shifted bands comparable to those of tBu-HBC, which contrasts with the broad red-shifted bands of [9]HBNG. This suggests a lower excimer-like contribution due to changes in the LUMO of oxa[9]HBNG (delocalized in the graphitized layers), as opposed to [9]HBNG whose LUMO is mainly localized in the helicene. Finally, the chiroptical properties of M- and P-oxa[9]HBNG have been studied by means of electronic circular dichroism, from which absorption dissymmetry factor (gabs) of +3.7·10−3 at 487 nm for M-oxa[9]HBNG, and −1.8·10−3 at 487 nm for P-oxa[9]HBNG, have been calculated. Resumen Resumen xiii RESUMEN Capítulo 1 La reacción de Scholl es clave en la síntesis de nanografenos moleculares, ya que da lugar a la grafitización de los poliarenos precursores. Las condiciones para llevar a cabo la reacción conllevan la utilización de ácidos de Lewis o ácidos de Brønsted en combinación con oxidantes como DDQ (2,3-dicloro-5,6-diciano-1,4-benzoquinona) y el mecanismo implica la formación de intermedios de tipo catión arenio o catión radical. Sin embargo, es difícil predecir el mecanismo operando y, en ocasiones, da lugar a compuestos inesperados como resultado de la formación de intermedios alternativos o reordenamientos. En este capítulo, se describe la síntesis inesperada de espironanografenos y como influyen los efectos electrónicos en el control sobre la reacción de Scholl. La reacción de Scholl sobre un poliareno basado en un núcleo central de antraceno da lugar a la formación de espirociclos que conectan las unidades grafitizadas con el antraceno central. Según revelaron los cálculos teóricos, esto ocurre debido a la formación de intermedios catión tritilo, muy estables, en las posiciones 9 y 10 del antraceno. Dependiendo de las condiciones de Scholl utilizadas se ha podido controlar el número de cierres de grafitización, describiendo así el primer ejemplo en el que la formación de espirociclos no interrumpe la grafitización. Persiguiendo el objetivo principal de sintetizar nanografenos helicoidalmente organizados formados por dos unidades grafitizadas (dibenzo[fg,ij]phenanthro- [9,10,1,2,3-pqrst]pentaphene, DBPP) conectadas a través de un antraceno central, se controló la reacción de Scholl mediante la introducción de ocho átomos de flúor. El carácter electroatractor de los átomos de flúor desestabiliza la formación del catión Resumen xiv tritilo en las posiciones 9,10 del antraceno, haciendo que la reacción evolucione por los intermedios de tipo catión arenio que dan lugar a la estructura grafitizada esperada. Además de la síntesis y caracterización estructural de los nuevos nanografenos moleculares, se ha llevado a cabo la caracterización de las propiedades optoelectrónicas. Las medidas de voltamperometría cíclica del espironanografeno mostraron valores de potenciales de oxidación y reducción más altos en valor absoluto (E1 1/2ox= 0.85 V, E1 1/2red= –2.56V) que los correspondientes al tBu-HBC E1 1/2ox= 0.67 V, E1 1/2red= –2.27 V), lo cual evidencia la interrupción de la conjugación en los carbonos espiránicos sp3 que conectan las unidades grafitizadas con el fragmento central de antraceno. Esta característica estructural también repercute sobre las propiedades ópticas de absorción y emisión, cuyas bandas están más desplazadas al azul, frente al tBu-HBC. Las propiedades electrónicas del nanografeno helicoidalmente organizado revelaron también la menor π-extensión de las unidades grafitizadas (DBPP), respecto al tBu- HBC, mostrando un primer potencial de oxidación E1 1/2ox= 0.87 V. Sin embargo, el carácter aceptor es mejor que en el tBu-HBC, ya que el primer potencial de reducción E1 1/2red= –1.69 V se ve altamente influenciado por la localización del LUMO en el fragmento electroaceptor de antraceno fluorado. Este comportamiento también se ve reflejado en el espectro de emisión, en el cual se observa un desplazamiento de las bandas al rojo. Curiosamente, esta estructura parece presentar una emisión dual, en la cual los dos fluoróforos (DBPPs y octafluoroantraceno) parecen dar lugar a las correspondientes bandas de emisión de manera independiente. Capítulo 2 La mayoría de las estrategias para la inducción de quiralidad en nanografenos implican la desviación de la planaridad mediante la introducción de defectos topológicos. Así, los ciclos no hexagonales y las unidades helicoidales son las características estructurales más estudiadas. Sin embargo, hasta la fecha, solo unos pocos ejemplos en la literatura representan la posibilidad de inducir quiralidad axial de tipo atropisomería, en la que la quiralidad surge de la restricción de giro de las unidades estructurales alrededor de un enlace sencillo. En este capítulo se describe una nueva estrategia para inducir atropisomería en nanografenos. Además, se estudia la dinámica molecular, directamente relacionada con la estabilidad de las propiedades quirópticas de los nanografenos dispuestos helicoidalmente. Los nanografenos descritos, formados por dos unidades grafitizadas (DBPP) conectadas a través de un fragmento central de tetrafluorobenzeno 1,4-disustituido, similares a la estructura basada en octafluoroantraceno obtenida en el capítulo anterior, se sintetizaron con el objetivo de simplificar el sistema para llevar a cabo un estudio de la quiralidad. Resumen xv La aproximación inicial consistía en la síntesis un nanografeno en el cual los sustituyentes en el DBPP son iguales y el fragmento central de benceno está sustituido por cuatro átomos de flúor. La posterior modificación de los sustituyentes daría lugar a atropisomería de sistemas terfenilo. Sin embargo, una vez obtenida la estructura en la que los DBPPs están sustituidos con cinco grupos tert-butilo y el benceno central con cuatro átomos de flúor en las posiciones 2,3,5,6, se observó la formación inesperada de dos enantiómeros en la estructura cristalina. Mediante cálculos DFT se pudo observar la formación de interacciones de tipo C−H···π entre las unidades de DBPP. Estas interacciones se dan como resultado de la imposibilidad de coplanaridad entre los DBPPs, y fijan la estructura con una organización helicoidal de las subunidades, dando lugar a dos imágenes especulares no superponibles. Los enantiómeros se observaron por HPLC quiral, sin embargo, no se pudieron aislar debido a su racemización a temperatura ambiente. Para el estudio de la barrera de isomerización se prepararon dos derivados en los cuales un grupo tert-butilo de cada DBPP se intercambió por un grupo OMe o un átomo de hidrógeno. Esta modificación implica la formación de dos diasterómeros casi degenerados diferenciables por RMN. A partir de una muestra enriquecida en uno de los diasterómeros se pudo monitorizar la evolución hasta el equilibrio (relación 1:1), lo cual permitió conocer el valor de la barrera de isomerización, Δ𝐺⧧= 24.6 kcal·mol−1 a 40 ºC, y del tiempo de vida media, τ1/2= 107 min. Además, con estos experimentos se pudo confirmar que el giro de 90º de la unidad de tetrafluorobenceno, es el motivo por el cual ocurre la isomerización. Adicionalmente, se estudiaron las propiedades electroquímicas por voltamperometría cíclica y por voltamperometría de onda cuadrada. Los nanografenos basados en tetrafluorobenceno son electrodadores y electroaceptores más débiles que el tBu- HBC, lo cual concuerda con la menor π-extensión de las unidades de DBPP y la débil conjugación a lo largo de la estructura. La misma tendencia se confirmó con las medidas de espectros de absorción y emisión, que muestran las bandas desplazadas hacia el azul respecto al tBu-HBC. Resumen xvi Capítulo 3 En este capítulo se describe la síntesis y caracterización de dos nanografenos bicapa helicoidales (HBNG) con diferente número de anillos orto-fusionados, dando lugar a dos helicenos de distinta longitud, [9] y [11]heliceno ([9]HBNG y [11]HBNG). Estas estructuras, junto con la publicada anteriormente por nuestro grupo de investigación, [10]HBNG (ACIE 2018), forman una familia cuyas características estructurales determinan la variación de las propiedades. En nanografenos moleculares es común la modulación de las propiedades mediante la variación de la extensión π o mediante la introducción de defectos en la red hexagonal (como se describe en el Capítulo 2). Sin embargo, en este trabajo se describe el impacto de una nueva característica estructural, el grado de solapamiento entre dos capas enfrentadas con interacciones π-π. Por tanto, en este capítulo se detalla el estudio sistemático de las propiedades optoelectrónicas y quirópticas de estos nanografenos. La longitud del heliceno determina el grado de solapamiento entre las capas grafitizadas fusionadas en los extremos, siendo [9]HBNG la más solapada (26 anillos intervienen en las interacciones π-π) y [11]HBNG (10 anillos intervienen en las interacciones π-π) la menos solapada. Además, en estas estructuras varía la extensión π con el número de anillos en el heliceno, siendo [9]HBNG la de menor extensión π y [11]HBNG la de mayor extensión π. Teniendo en cuenta estas características estructurales, se pudo estudiar y comprender el impacto del grado de solapamiento sobre las propiedades. El carácter dador más fuerte lo presenta el nanografeno de menor extensión π, [9]HBNG. Este comportamiento se atribuye a un efecto de valencia mixta mediante el cual el catión y el catión radical (especies oxidadas) se estabilizan entre las capas grafitizadas. Además, esta estructura presenta las bandas de emisión más desplazadas al rojo, evidenciando la comunicación de las capas a través del espacio por interacciones π-π. Curiosamente, la alineación de los vectores eléctrico y magnético hacen que [9]HBNG sea el nanografeno con mejores propiedades quirópticas, con un factor de Mayor solapamiento Mayor extensión π Caracter dador más fuerte [9]HBNG [10]HBNG [11]HBNG ACIE 2018 Resumen xvii absorción y emisión de luz circularmente polarizada (gabs y glum) de 3.6·10−2, uno de los más altos descritos hasta ahora para estructuras de carbono. Capítulo 4 El creciente interés en nanografenos bicapa helicoidales debido principalmente a las características singulares que aporta el solapamiento π-π y a las destacables propiedades quirópticas (Capítulo 3), hace necesaria la búsqueda de nuevas estrategias para mejorar la accesibilidad sintética. En este capítulo, se describe una estrategia sintética de nanografenos bicapa quirales fácilmente escalable evitando la costosa separación de enantiómeros por HPLC quiral. La síntesis parte del compuesto 7-bromo-2-naftol, comercialmente disponible, cuya reactividad permite la obtención de derivados de BINOL 7,7’-disustituido. La estabilidad configuracional del BINOL (con quiralidad axial) y la reactividad de los grupos naftol, han permitido la esterificación con cloruro de canforsulfonilo enantioméricamente puro para la resolución química de los enantiómeros del poliareno basado en BINOL. Posteriormente se ha realizado la reacción de Scholl, en la cual ocurre la ciclodeshidrogenación (grafitización) y la ciclodeshidratación enantioespecífica de los grupos naftol dando lugar al oxa[9]HBNG con transferencia de la quiralidad axial a quiralidad helicoidal. Además, se ha llevado a cabo la caracterización de las propiedades optoelectrónicas y quirópticas. Mediante voltamperometría cíclica se observó la influencia del solapamiento (22 anillos intervienen en las interacciones π-π intramoleculares), la estructura presenta un carácter dador ligeramente inferior, E1 1/2ox= 0.37 V, a la bicapa Resumen xviii [9]HBNG (26 anillos intervienen en las interacciones π-π intramoleculares), E1 1/2ox= 0.35 V. Sin embargo, la variación más destacada se observa en el carácter aceptor, la presencia del átomo de oxígeno en el heliceno hace que el LUMO se localice en las capas grafitizadas, aumentando el potencial de reducción, E1 1/2red= −2.27 V, a valores similares a los de tBu-HBC, E1 1/2red= −2.24 V. Además, el espectro de emisión del oxa[9]HBNG muestra bandas bien estructuradas desplazadas al azul y comparables a las del tBu-HBC, lo cual contrasta con las bandas anchas desplazadas al rojo del [9]HBNG. Esto sugiere una menor contribución de tipo excímero debido a los cambios en el LUMO del oxa[9]HBNG (deslocalizado en las capas grafitizadas), frente a [9]HBNG cuyo LUMO se localiza principalmente en el heliceno. Finalmente, las propiedades quirópticas de los enantiómeros puros M-oxa[9]HBNG y P- oxa[9]HBNG se han estudiado por dicroísmo circular, de donde se han obtenido unos factores de disimetría de absorción (gabs) de +3.7·10−3 a 487 nm para M- oxa[9]HBNG, y −1.8·10−3 a 487 nm para P-oxa[9]HBNG. Introduction Introduction 1 INTRODUCTION The first 2D atomic crystal: a honeycomb patterned carbon layer “Graphene is the name given to a single layer of carbon atoms densely packed into a benzene-ring structure”.1 This is the definition provided by Konstantin Novoselov and Andre Geim in their seminal work published in Science 2004, in which they isolate graphene for the first time. Until that moment, graphene had been broadly explored as a theoretical model2,3,4 and was believed to be unstable in the free state with respect to the formation of curved structures as fullerenes or nanotubes. Defying the predictions against its existence, monolayer graphene was unexpectedly obtained in a way that was as simple as it was ingenious. Using adhesive tape, it was possible to exfoliate graphene samples from small fragments of highly oriented pyrolytic graphite. The first AFM image of single-layer graphene and its chemical structure are represented in Figure 1. Along with the discovery of graphene, the striking electronic properties positioned this material as the best possible “metal” for metallic transistor applications, offering also ballistic transport. In 2005, it was confirmed that the charge carriers in graphene, that mimic relativistic particles with zero rest mass having an effective speed of light, are massless Dirac fermions.5 Thus, from the electronic properties point of view, graphene is a zero-gap semiconductor, Figure 1, right. Figure 1. First AFM image of single-layer graphene (left, from ref 1), chemical structure of graphene (middle), and band structure of graphene with Dirac cones located on a hexagonal plane (right). This first two-dimensional atomic crystal set a milestone in materials and condensed matter science. Only three years after the discovery, in 2007, Geim and Novoselov reported a compilation of the properties and potential applications already observed, 1 K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva, A. A. Fisov, Science 2004, 306, 666. 2 P. R. Wallace, Phys. Rev. 1947, 71, 622. 3 J. C. Slonczewski, Phys. Rev. 1958, 109, 272. 4 G. W. Semenoff, Phys. Rev. Lett. 1984, 53, 2449. 5 K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, M. I. Katsnelson, I. V. Grigorieva, S. V. Dubonos, A. A. Firsov, Nature 2005, 438, 197. Introduction 2 The rise of graphene,6 thus providing clear evidence of the rapid interest of the scientific community in this two-dimensional material. The accessible procedures to obtain high-quality graphene allowed the fast progress of the research. Before 2012 the study of the material parameters had already shown outstanding properties. For instance, the electron mobility at room-temperature is of 2.5·10−5 cm2V−1s−1,7 which exceeds the theoretical limit of 2·10−5 cm2V−1s−1.8 The Yong’s modulus, which describes the materials stiffness, is of 1 TPa and the tensile strength is of 130 GPa.9 The extreme strength is particularly striking in view of the light weigh (0.77 mg/m2). Furthermore, it presents high thermal conductivity,10 complete impermeability to gases,11 transparency, and the ability to support very high densities of electric current (a million times higher than copper).12 This combination of properties made graphene a promising material for applications in disruptive technologies.13 However, these properties and therefore, the potential applications depend on the quality and size of the crystals, which are strongly affected by the production method. Nowadays there are several methods for the mass-production of graphene,14 which allows to control the size, quality, and prize, for particular applications.15 The methodologies can be classified in two main approaches, top-down and bottom-up. Top-down methods consisting of the exfoliation of graphite, is a straightforward and widely used method that can be mechanical, chemical16 or electrochemical.17 Conversely, the bottom-up approach involves the formation of graphene from smaller materials at the atomic or nanometer scale. These methods include chemical vapor 6 A. Geim, K. Novoselov, Nat. Mater. 2007, 6, 183. 7 A. S. Mayorov, Nano Lett. 2011, 11, 2396. 8 S. V. Morozov, Phys. Rev. Lett. 2008, 100, 016602. 9 C. Lee, X. D. Wei, J. W. Kysar, J. Hone, Science 2008, 321, 385. 10 A. A. Balandin, Nat. Mater. 2011, 10, 569. 11 J. S. Bunch, S. S. Verbridge, J. S. Alden, A. M. van der Zande, J. M. Parpia, H. G. Craighead, P. L. McEuen, Nano Lett. 2008, 8, 2458. 12 J. Moser, A. Barreiro, A. Bachtold, Appl. Phys. Lett. 2007, 91, 163513. 13 A. K. Worku, D. W. Ayele, Results Chem. 2023, 5, 100971. 14 A. Gutiérrez-Cruz, A. R. Ruiz-Hernández, J. F. Vega-Clemente, D. G. Luna-Gazcón, J. Campos- Delgado, J. Mater. Sci. 2022, 57, 14543. 15 K. S. Novoselov, V. I. Fal’ko, L. Colombo, P. R. Gellert, M. G. Schwab, K. Kim, Nature 2012, 490, 192. 16 K. Parvez, S. Yang, X. Feng, K. Müllen, Synth. Met. 2015, 210, 123. 17 N. Liu, F. Luo, H. Wu, Y. Liu, C. Zhang, J. Chen, Adv. Funct. Mater. 2008, 18, 1518. Introduction 3 deposition,18,19 epitaxial growth,20 laser-induced,21 laser ablation22 and electric discharge “flash” synthesis.23 Functionalization of graphene Alternatively, one of the most exploited properties of graphene is its chemical reactivity. The functionalization of the honeycomb backbone is a widely extended method for the chemical preparation of graphene derivatives.24 The treatment of graphite with oxidative reagents, commonly acids, entails the oxygenation of the layers in the borders, as well as in the basal planes, providing the commonly known graphene oxide after exfoliation. This material exhibits a mixed sp2−sp3 carbon system, and due to its non-conductivity and relative hydrophobicity, it has been widely applied in medicine.25,26 However, for further applications, it is required the restoring of the π network to recover the electric conductivity. The “reparation” of the carbon network requires the removal of the oxygen-containing functional groups in the layers. This reduction process to restore the honeycomb lattice has been achieved by different chemical and non-chemical methods,14 and the layer obtained is known as reduced graphene oxide. Although the number of remaining defects is important, this material has been used in composites, thermal dissipation films and conducting additives.27 To sum up, the possibilities to obtain graphene are currently wide and technically diverse. However, it is important to note that the production methodology is directly related to the quality of the material and its properties, which determines its potential applications. Since the early years of graphene's rise as a tangible and manipulable material, its discoverers quickly realized its potential applications, especially in electronics. Although, they warned: “despite the reigning optimism about graphene-based electronics, ‘graphenium’ microprocessors are unlikely to appear for the next 20 18 K. S. Kim, Y. Zhao, H. Jang, S. Y. Lee, J. M. Kim, K. S. Kim, J.-H. Ahn, P. Kim, J.-Y. Choi, B. H. Hong, Nature 2009, 457, 706. 19 X. Li, W. C. J. An, S. K. J. Nah, D. Yang, R. Piner, A. Velamakanni, I. Jung, E. Tutuc, S. K. Banerjee, L. Colombo, R. S. Ruoff, Science 2009, 324, 1312. 20 C. Berger, Z. Song, T. Li, X. Li, A. Y. Ogbazghi, R. Feng, Z. Dai, A. N. Marchenkov, E. H. Conrad, P. N. First, W. A. De Heer, J. Phys. Chem. B 2004, 108, 19912. 21 J. Lin, Z. Peng, Y. Liu, F. Ruiz-Zepeda, R. Ye, E. L. G. Samuel, M. J. Yacaman, B. I. Yakobson, J. M. Tour, Nat. Commun. 2014, 5, 5714. 22 M. Qian, Y. S. Zhou, Y. Gao, J. B. Park, T. Feng, S. M. Huang, Z. Sun, L. Jiang, Y. F. Lu, Appl. Phys. Lett. 2011, 98, 173108. 23 D. X. Luong, K. V. Bets, W. A. Algozeeb, M. G. Stanford, C. Kittrell, W. Chen, R. V. Salvatierra, M. Ren, E. A. McHugh, P. A. Advincula, Z. Wang, M. Bhatt, H. Guo, V. Mancevski, R. Shahsavari, B. I. Yakobson, J. M. Tour, Nature 2020, 577, 647. 24 A. Adetayo, D. Runsewe, Open J. Compos. Mater. 2019, 9, 207. 25 Y. Zhu, S. Murali, W. Cai, X. Li, J. W. Suk, J.R. Potts, R.S. Ruoff, Adv. Mater. 2010, 22, 3906. 26 N. Panwar, A. M. Soehartono, K. K. Chan, S. Zeng, G. Xu, J. Qu, P. Coquet, K.-T. Yong, X. Chen, Chem. Rev. 2019, 119, 9559. 27 L. Lin, H. Peng, Z. Liu, Nat. Mater. 2019, 18, 520. Introduction 4 years”.5 Currently, graphene has been applied to a wide range of fields, including electronics, photonics and optoelectronics,28,29,30 fuel cells,31 energy storage,32 artificial intelligence,33 biomedicine,34 and even cultural heritage.35 Thus, in a situation in which graphene is produced in a several thousand tons/year36 scale in the present times, it is worth to question what else can be done. Trends and Recent Developments Bilayer graphene: twist to superconductivity Paradoxical though it may seem, one of the current challenges and hot topics in graphene is the overlapping of a fixed number of graphene layers by controlling the relative twist angle between them. Interestingly, this growing interest, especially in graphene bilayers, started in the same year in which Andre Geim and Konstantin Novoselov were awarded with Nobel Prize in Physics (2010) "for groundbreaking experiments regarding the two-dimensional material graphene" (Figure 2). Figure 2. The same year that Geim and Novoselov received the Nobel Prize in Physics for graphene, new possible phases of matter were expected from twisted bilayer graphene. In 2010, Eva Y. Andrei and coworkers reported the observation of van Hove singularities in overlapping 2D crystals of graphene with a 1.79º relative twist between the layers (Figure 3).37 This observation resulted very intriguing considering that new phases of matter can arise near the van Hove singularities (saddle point in the electronic band structure leading to a divergence in the density of states, (Figure 3).38 Thus, indicating the possibility of modulating graphene properties without changing its chemical structure, just controlling the number of layers and their relative 28 A. N. Grigorenko, M. Polini, K. S. Novoselov, Nat. Photonics 2012, 6, 749. 29 J. Wu, H. Lin, D. J. Moss, K. P. Loh, B. Jia, Nat. Rev. Chem. 2023, 7, 162. 30 F. Bonaccorso, Z. Sun, T. Hasan, A. C. Ferrari, Nat. Photonics 2010, 4, 611. 31 H. Su, Y. H. Hu, Energy Sci. Eng. 2021, 9, 958. 32 F. Bonaccorso, L. Colombo, G. Yu, M. Stoller, V. Tozzini, A. C. Ferrari, R. S. Ruoff, V. Pellegrini, Science 2015, 347, 1246501. 33 M. Huang, Z. Li, H. Zhu, Adv. Intell. Syst. 2022, 4, 2200077. 34 J. Li, H. Zeng, Z. Zeng, Y. Zeng, T. Xie, ACS Biomater. Sci. Eng. 2021, 7, 5363. 35 E. Galvagno, E. Tartaglia, M. Stratigaki, C. Tossi, L. Marasco, F. Menegazzo, C. Zanardi, F. Omenetto, C. Coletti, A. Traviglia, M. Moglianetti, Adv. Funct. Mater. 2024, 2313043. 36 M. T. U. Safian, K. Umar, M. N. Mohamad Ibrahim, J. Clean. Prod. 2021, 318, 128603. 37 G. Li, A. Luican, J. M. B. Lopes dos Santos, A. H. Castro Neto, A. Reina, J. Kong, E. Y. Andrei, Nat. Phys. 2010, 6, 109. 38 L. Van Hove, Phys. Rev. 1953, 89, 1189. Nobel Prize in Physics 2010 Geim and Novoselov Van Hove singularities, twisted bilayer graphene Introduction 5 twist angle. These early observations were the starting point for the development of new metamaterials. Figure 3. Saddle point (sp) between the two Dirac points (from ref 37) in 1.79º twist bilayer. The stacking of two or more atomically thin materials with a relative twist angle or a lattice mismatch leads to so-called moiré quantum materials. In 2018, Jarillo-Herrero and coworkers described the results expected from the previous observations, a new phase of matter in twist bilayer graphene, unconventional superconductivity.39 The interlayer interaction between two stack graphene layers induces the hybridization of their bands, entailing modifications in the low-energy band structure depending on the stacking order (AA or AB, Figure 4). Additionally, the relative twist angle between the layers modulates the hexagonal moiré patterns, consisting of alternating AA and AB stacked regions, in the resulting superlattice, Figure 3 left. The precise control over the twist angles revealed that at the particular twist angle of 1.1º, termed as the first magic angle, the phase diagram of twisted bilayer graphene is similar to that of high-temperature superconductors. Figure 4. Representation of the stacking possibilities between graphene layers, AA (left), AB (right). In view of these results and the properties reported so far,40 it is expected that twisted bilayer and multilayer graphene will follow the successful path of monolayer graphene, with new applications where carbon materials were not sufficiently 39 Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi, E. Kaxiras, P. Jarillo-Herrero, Nature 2018, 556, 43. 40 E. Y. Andrei, D. K. Efetov, P. Jarillo- Herrero, A. H. MacDonald, K. Fai Mak, T. Senthil, E. Tutuc, A. Yazdani, A. F. Young, Nat. Rev. Mater. 2021, 6, 201. A A A B Introduction 6 competitive. It is striking that a simple layer with a honeycomb pattern and the thickness of a carbon atom, whose existence was so often questioned, has revolutionized the science of materials and condensed matter, allowing the development of disruptive technologies already introduced in the society 20 years after its discovery. Graphene molecules: Nanographenes Polycyclic aromatic hydrocarbons have attracted the interest of the scientific community probably since Michael Faraday discovered benzene in 1825. The multidirectional evolution of sp2-hybridised carbon materials ranges from the development of simple molecules to fullerenes (discovered in 1985 by Kroto, Curl and Smalley), carbon nanotubes (discovered in 1991, by Iijima and Ichihashi) and graphene. The knowledge gained in organic reactions and new characterization techniques, has enabled the continuous evolution of carbon-based materials. It is therefore not so surprising that organic stepwise synthesis is one of the most powerful tools for producing graphene molecular fragments. However, it results interesting that the grater development of these nanometric-sized fragments, with honeycomb pattern, took place in parallel to the emergence of single-layered graphene and its derivatives. The smallest and most frequently investigated graphene molecule, hexa-peri- hexabenzocoronene (HBC), was described for the first time by Halleux in 1958.41 This flat structure formed by 42 carbon atoms was for a long time the largest fully characterized polyaromatic hydrocarbon (Figure 5). However, the development of different methodologies based on stepwise organic synthesis has made possible the monodisperse preparation of polycyclic aromatic hydrocarbons with different sizes and shapes.42,43 For instance, in 2002, Klaus Müllen et al., pioneers in the preparation of molecular graphene fragments, reported one of the largest structures described so far (the structure is formed by 222 carbon atoms and has a diameter of 3 nm), demonstrating the potential of organic chemistry for obtaining large graphene molecules (Figure 5).44 41 A. Halleux, R. H. Martin, G. S. D. King, Helv. Chim. Acta 1958, 41, 1177. 42 L. Chen, Y. Hernandez, X. Feng, K. Müllen, Angew. Chem. Int. Ed. 2012, 51, 7640. 43 J. Liu, X. Feng, Angew. Chem. Int. Ed. 2020, 59, 23386. 44 C. D. Simpson, J. D. Brand, A. J. Berresheim, L. Przybilla, H. J. Räder, K. Müllen, Chem. Eur. J. 2002, 8, 1424. Introduction 7 Figure 5. From benzene and small polyaromatic hydrocarbons to nanographenes. In 2012 Müllen and coworkers defined as nanographenes those graphene fragments ranging from 1 to 100 nm in size.42 The last decade has marked the boom of nanographenes, with a large number of structures reported in the literature.45 From the first flat molecules to increasingly sophisticated structures that have demonstrated surprising optoelectronic properties, as it will be described in the following chapters. The main reason for this still growing interest, is that nanographenes have proved to be a suitable alternative in those applications where graphene has more limitations. The zero-gap semiconducting nature of graphene limits its application on optoelectronics. In contrast, the quantum confinement of the electrons in smaller molecular structures, opens the gap between the conduction band and the valence band, thus broadening the range of applications in sensing, electronics, or photovoltaics.46 Several review publications on the synthesis and modulation of nanographenes properties are currently available.45,47,48,49 Therefore, as with graphene, one wonders what’s next. Surprisingly enough, it seems that graphene and its molecular derivatives are progressing along the same lines. In the same year that Jarillo-Herrero described the superconductivity properties of bilayer graphene (2018),39 our research group described the first molecular structure with two covalently connected overlapping layers. This marked the beginning of helical bilayer nanographenes, whose rise we are currently witnessing, mainly due to their remarkable chiroptical properties, as it will be described and discussed in the following chapters. 45 Z. Liu, S. Fu, X. Liu, A. Narita, P. Samori, M. Bonn, H. I. Wang, Adv. Sci. 2022, 9, 2106055. 46 Y. Gu, Z. Qiu, K. Müllen, J. Am. Chem. Soc. 2022, 144, 11499. 47 A. Narita, X.-Y. Wang, X. Feng, K. Müllen, Chem. Soc. Rev. 2015, 44, 6616. 48 S. Song, J. Su, M. Telychko, J. Li, G. Li, Y. Li, C. Su, J. Wu, J. Lu, Chem. Soc. Rev. 2021, 50, 3238. 49 Y. Segawa, H. Ito, K. Itami, Nat. Rev. Mater. 2016, 1, 15002. Chapter 1 Chapter 1 11 1 Chapter 1 1.1 Introduction to the Synthesis of Nanographenes The bottom-up approach for the synthesis of nanographenes allows the monodisperse preparation of graphene molecular fragments with atomic control. Considering hexa- peri-hexabenzocoronene (HBC) the smallest nanographene, the general synthetic methodology, consists in the synthesis of hexaphenylbenzene (1.1) and the subsequent formation of multiple carbon-carbon bonds to obtain the graphitized nanographene, HBC (1.2), Figure 1.1. Figure1.1. General synthetic methodology of molecular nanographenes represented with the smallest nanographene, HBC. 1.1.1 Key Reactions for the π-Extension In the first half of the 20th century, Roland Scholl and Eric Clar were pioneers to report the synthesis and characterization of polyaromatic hydrocarbons.50 Since then, the extensive development of organic chemistry reactions has provided powerful tools for the design, synthesis, and characterization of advanced novel aromatic carbon-based structures. Although numerous strategies for the synthesis of nanographenes can be found in the literature, a few key reactions can be identified. The previous stages of precursor synthesis require the generation of polyarene-like structures, in which the π-extension is increased, and the rings are placed in suitable arrangements for subsequent graphitization. The most widely and efficiently proven reactions for the π-extension are Diels-Alder cycloaddition, metal-mediated cyclotrimerization of alkynes and metal-catalyzed cross couplings. In 2000, Müllen and coworkers reported the first synthesis of hexaphenylbenzene derivatives by a Diels-Alder reaction between tetraphenylcyclopentadienones 1.3 as dienes and aromatic alkynes 1.4 as dienophiles, to achieve the synthesis of 50 (a) R. Scholl, C. Seer, R. Weitzenbök, Chem. Ber. 1910, 43, 2202. (b) R. Scholl, C. Seer, Liebigs Ann. Chem. 1912, 394, 111. (c) R. Scholl, C. Seer, Chem. Ber. 1922, 55, 330. (d) E. Clar, D. G. Stewart, J. Am. Chem. Soc. 1953, 75, 2667. (e) E. Clar, W. Schmidt, Tetrahedron 1979, 35, 2673. Chapter 1 12 hexaarylbenzenes 1.5 with different substitution patterns.51 By following this strategy, the substituents can be incorporated, as shown in Scheme 1.1, through the functionalization of the cyclopentadienone and the acetylene. Scheme 1.1. Synthetic strategy for polyarenes by Diels-Alder cycloaddition. (i) Synthesis of diarylacetylene, commonly by Sonogashira coupling. (ii) Aldolic condensation yields cyclopentadienones (or cyclones). (iii) Diels-Alder cycloaddition to obtain polyarenes. The [4+2] cycloaddition entails the formation of a new benzene ring decorated with six functionalized aryl substituents. The reaction occurs with carbon monoxide extrusion resulting from a chelotropic elimination once the six-member ring is formed, leading to the aromatization of the central benzene ring. This methodology has been widely used in the synthesis of nanographenes with different shapes and sizes. The possibilities to vary the cyclopentadienone as well as the alkyne, and their substitution pattern, allows to obtain a great variety of structures. For instance, Wang and coworkers reported the twistacene nanographene52 1.6 synthesized from a two-fold Diels-Alder cycloaddition between the p-phenylene- linked cyclopentadienone dimer 1.7 and bis(4-tert-butylphenyl)acetylene (1.8) that yielded the polyarene 1.9 showed in Scheme 1.2. 51 S. Ito, M. Wehmeier, J. D. Brand, C. Kübel, R. Epsch, J. P. Rabe, K. Müllen, Chem. Eur. J. 2000, 6, 4327. 52 S. Ma, J. Gu, C. Lin, Z. Luo, Y. Zhu, J. Wang, J. Am. Chem. Soc. 2020, 142, 16887. Chapter 1 13 Scheme 1.2. Desing of cyclopentadienones to perform the Diels-Alder cycloaddition. Following the [4+2] approach, Müllen et al. reported the synthesis of large graphene nanoribbon 1.10.53 Starting from the π-extended building block with dual functionalization 1.11, the cyclopentadienone and the terminal alkyne (Scheme 1.3), the Diels-Alder cycloaddition leads to the polyarene polymer 1.12. 53 A. Narita, I. A. Verzhbitskiy, W. Frederickx, K. S. Mali, S. A. Jensen, M. R. Hansen, M. Bonn, S. De Feyter, C. Casiraghi, X. Feng, K. Müllen, ACS Nano 2014, 8, 11622. Chapter 1 14 Scheme 1.3. Polymerization by Diels-Alder cycloaddition to prepare graphene nanoribbons. The possibility of modifying the alkyne precursor has also been explored. For example, Martín et al. reported the synthesis of curved nanographene 1.13 starting from cyclooctatetraene with four fused thiophenes substituted with four 4-tert- butylphenylalkynyl groups 1.14 (Scheme 1.4). The polyarene 1.15 is obtained after performing a fourfold Diels-Alder reaction with the corresponding cyclopentadienone.54 54 J. Urieta-Mora, M. Krug, W. Alex, J. Perles, I. Fernández, A. Molina-Ontoria, D. M. Guldi, N. Martín, J. Am. Chem. Soc. 2020, 142, 4162. Chapter 1 15 Scheme 1.4. Synthesis of four-thiophene-fused cyclooctatetraene-based polyarene by Diels- Alder, followed by graphitization to afford a curved nanographene. An alternative to the [4+2] cycloaddition approach is the metal-mediated cyclotrimerization of aromatic alkynes 1.16. This reaction has proved to be highly efficient when the diarylacetylenes involved are symmetrically substituted. As shown in Scheme 1.5, a new benzene ring is formed by joining the aryl substituents coming from the acetylene derivative resulting in a polyarene 1.17. Scheme 1.5. Synthesis of polyarenes by metal-mediated cyclotrimerization of alkynes. Chapter 1 16 Although the reaction efficiency is usually higher when the alkyne precursor is symmetrically substituted, Campaña and coworkers reported the large triskelion- shaped nanographene55 1.18 in which the alkyne 1.19 is asymmetrically substituted with a tert-butylphenyl group and a hexabenzocoronene derivative at each side. However, the reaction led to the preferential formation of the 1,3,5-substituted regioisomer, the polyarene 1.20 (Scheme 1.6). Scheme 1.6. Synthesis of polyarene by cyclotrimerization of asymmetric alkynes followed by graphitization to afford a triskelion-shaped nanographene. The third most widely used methodology for the design of polyarenes, e.g. 1.21, consists in metal-catalyzed cross-couplings. This alternative approach allows changing features as the geometry, the substitution pattern, or the π-extension degree, as shown in Scheme 1.7. The variety of cross-coupling reactions available, Suzuki, Stille, Yamamoto, Heck... provides great versatility for the modification of the structures. 55 C. M. Cruz, I. R. Márquez, S. Castro-Fernández, J. M. Cuerva, E. Maçôas, A. G. Campaña, Angew. Chem. Int. Ed. 2019, 58, 8068. Chapter 1 17 Scheme 1.7. Synthesis of polyarenes by metal-mediated cross-coupling reactions. Following this methodology, Itami and coworkers reported the synthesis of the highly curved nanographene 1.22 from different polyarene precursors.56 The introduction of 4-tert-butylphenyl or biphenyl substituents in a corannulene moiety (1.23 and 1.24, respectively) by cross-coupling reactions influence the following graphitization step, as different bonds remained to be formed (Scheme 1.8). Scheme 1.8. Synthesis of two polyarenes through metal-mediated cross-couplings leading to a curved nanographene after oxidation. 56 K. Kawasumi, Q. Zhang, Y. Segawa, L. T. Scott, K. Itami, Nat. Chem. 2013, 5, 739. Chapter 1 18 1.1.2 Graphitization The key step in the synthesis of nanographenes is the graphitization of the polyarene precursor. To this end, different reactions can be found in the literature. For instance, it can be accomplished by intramolecular catalyzed couplings (Scheme 1.9a),57 C−F bond activation, methodology developed by Amsharov et al. (Scheme 1.9b),58 photochemical cyclization (Scheme 1.9c),59 or flash vacuum pyrolysis, as described by Scott et al.60 to achieve the formation of a highly curved π-extended corannulene (Scheme 1.9d). These reactions require the previous functionalization of the polyarene precursors, commonly with halogens. Scheme 1.9. Graphitization reactions starting from functionalized polyarenes. (a) Intramolecular metal-catalyzed cross-coupling. (b) C−F bond activation. (c) Photochemical cyclization. (d) Flash vacuum pyrolysis. Nevertheless, the most widely used reactions in the synthesis of molecular nanographenes are those that do not require the introduction of reactive groups for the subsequent graphitization. In this context, electrochemical and on-surface reactions are rapidly developing methodologies. Kong et al. reported an electrochemical cyclodehydrogenation methodology that requires catalytic equivalents of DDQ (2,3- dichloro-5,6-dicyano-1,4-benzoquinone) (Scheme 1.10a).61 On-surface reactions have proven to be suitable for the graphitization of planar structures. The interaction 57 X.-S. Zhang, Y.-Y. Huang, J. Zhang, W. Meng, Q. Peng, R. Kong, Z. Xiao, J. Liu, M. Huang, Y. Yi, L. Chen, Q. Fan, G. Lin, Z. Liu, G. Zhang, L. Jiang, D. Zhang, Angew. Chem. Int. Ed. 2020, 59, 3529. 58 V. Akhmetov, M. Feofanov, O. Papaianina, S. Troyanov, K. Amsharov, Chem. Eur. J. 2019, 25, 11609. 59 M. Daigle, A. Picard-Lafond, E. Soligo, J.-F. Morin, Angew. Chem. Int. Ed. 2016, 55, 2042. 60 L. T. Scott, E. Jackson, Q. Zhang, B. D. Steinberg, M. Bancu, B. Li, J. Am. Chem. Soc. 2012, 134, 107. 61 W.-J. Kong, L. H. Finger, J. C. A. Oliveira, L. Ackermann, Angew. Chem. Int. Ed. 2019, 58, 6342. Chapter 1 19 of the starting molecules with a metal surface after temperature treatment induce graphitization (Scheme 1.10b).62 Scheme 1.10. Graphitization reactions starting from unfunctionalized polyarenes. (a) Electrochemical cyclodehydrogenation. (b) On-surface Au(111) graphitization. However, most molecular nanographenes reported in the last decade have been obtained by conventional organic chemistry in solution. For this purpose, the preferred graphitization reaction is the well-known Scholl oxidation. 1.1.3 Scholl Cyclodehydrogenation The most extended graphitization reaction for molecular nanographenes is the Scholl oxidation (Roland Scholl, 1910).63,64 This reaction is a cyclodehydrogenation that occurs by double aryl C−H activation leading to the formation of multiple C−C covalent bonds.65,66 The reagents to perform this oxidation are Lewis acids (commonly AlCl3, MoCl5 and FeCl3) or oxidants in presence of Brønsted acids (as 2,3-dichloro- 5,6-dicyano-1,4-benzoquinone, DDQ, or para-chloranil in combination with methane sulfonic acid or trifluoromethanesulfonic acid). The Scholl cyclodehydrogenation mechanism has been investigated on multiple occasions, both experimental and theoretically. However, it remains controversial. The most accepted pathways involve the formation of arenium cation or radical-cation intermediates, but discerning the operating mechanism is usually unpredictable. One of the major difficulties for the in-depth evaluation to understand the mechanism is 62 B. Cirera, A. Sánchez-Grande, B. de la Torre, J. Santos, S. Edalatmanesh, E. Rodríguez-Sánchez, K. Lauwaet, B. Mallada-Faes, R. Zbořil, R. Miranda, O. Gröning, P. Jelínek, N. Martín, D. Écija, Nat. Nanotechnol. 2020, 15, 437. 63 (a) R. Scholl, J. Mansfeld, Ber. Dtsch. Chem. Ges. 1910, 43, 1734. (b) R. Scholl, C. Seer, Justus Liebigs, Ann. Chem. 1912, 394, 111. 64 R. S. Jassas, E. U. Mughal, A. Sadiq, R. I. Alsantali, M. M. Al-Rooqi, N. Naeem, Z. Moussa, S. A. Ahmed, RSC Adv. 2021, 11, 32158. 65 M. Grzybowski, K. Skonieczny, H. Butenschçn, D. T. Gryko, Angew. Chem. Int. Ed. 2013, 52, 9900. 66 M. Grzybowski, B. Sadowski, H. Butenschçn, D. T. Gryko, Angew. Chem. Int. Ed. 2020, 59, 2998. Chapter 1 20 that the Lewis acids commonly used are also oxidizing agents, and the alternative requires the combination of oxidants and protic acids. The arenium cation intermediates pathway67,68 is proposed to evolve as shown in Scheme 1.11. It starts with the protonation of aryl species, e. g. 1.25, forming the intermediate Int 1, an electrophilic σ-complex in case of performing the reaction with a Lewis acid, or a cation when using a Brønsted acid. The nucleophilic attack from the neighboring benzene ring leads to the formation of a new C−C bond, Int 2. The first proton elimination leads to the rearomatization of the first benzene ring, Int 3. The subsequent protons and electrons elimination regenerates the aromaticity in the system, giving the product 1.26. Scheme 1.11. Scholl reaction mechanism via arenium cation intermediates. Alternatively, the reaction may occur through the formation of radical-cation intermediates,69,70,71 such as Int 4 resulting from a single-electron oxidation of 1.27 (Scheme 1.12). The neighboring nucleophilic phenyl group attacks the radical cation to form the new bond Int 5, the subsequent oxidation and deprotonation yields the corresponding arenium cation intermediate Int 6, which after deprotonation leads to the aromatized compound 1.28. Scheme 1.12. Scholl reaction mechanism via radical-cation intermediates. The difficulty in identifying the mechanism operating on a given substrate is still challenging and under continuous study, especially in nanographenes chemistry. The formation of unexpected products because of ring migrations, rearrangements, transpositions or alternative cyclizations remains mostly unpredictable.72,73 It seems to be serendipitous and dependent of multiple factors, such as the molecular structure of the precursor, the steric hindrance and strain, the aromaticity, the reaction 67 G. Baddeley, J. Kenner, J. Chem. Soc. 1935, 303. 68 G. Baddeley, J. Chem. Soc. 1950, 994. 69 J. Kenner, J. Soc. Chem. Ind. 1933, 42, 470. 70 L. Zhai, R. Shukla, S. H. Wadumethrige, W. Rathore, J. Org. Chem. 2010, 75, 4748. 71 L. Zhai, R. Shukla, R. Rathore, Org. Lett. 2009, 11, 3474. 72 N. Ponugoti, V. Parthasarathy, Chem. Eur. J. 2022, 28, e202103530. 73 Y. Zhang, S. H. Pun, Q. Miao, Chem. Rev. 2022, 122, 14554. Chapter 1 21 conditions, etc. For example, Wu et al. demonstrated that the use of different Scholl reaction conditions on the same polyarene 1.29 can proceed through different mechanisms and yield different products. As concluded by DFT calculations, radical cation intermediates lead to nanographene 1.30 bearing hexagons¸ and cation intermediates lead to nanographene 1.31 bearing two octagonal rings (Scheme 1.13).74 Scheme 1.13. Scholl reaction conditions control the mechanism and the consequent nanographene structure. 1.1.3.1 Rearrangements in the Scholl Reaction Aryl migrations and isomerization reactions during the Scholl oxidation have been widely observed and have proven to be involved in most of the rearrangements occurring in presence of the typical oxidants and acids. Sigura and coworkers reported an aryl migration from 1.32 that occurs through the formation of radical cation intermediates and involves an attack by MeOH as nucleophile, evolving to the ketone containing product 1.33, without observing the formation of the expected tetrabenzocoronene structure 1.34 or alternative graphitization (Scheme 1.14).75 74 Y. Zou, Y. Han, S. Wu, X. Hou, C. H. E. Chow, J. Wu, Angew. Chem., Int. Ed. 2021, 60, 17654. 75 F. Ko, K. Hirabayashi, T. Shimizu, K.-I. Sugiura, Tetrahedron 2018, 59, 4251. Chapter 1 22 Scheme 1.14. Aryl migration under Scholl reaction conditions involving the incorporation of a carbonyl group from methanol. However, rearrangements can lead to ring cyclizations. The formation of five- membered rings together with a rearrangement was reported by Müllen and coworkers (Scheme 1.15). Performing the Scholl reaction over 1.35 to obtain the six-membered π-extended molecule 1.36 yielded the unexpected formation of two five-membered rings and the double 1,2-aryl shift of the tert-butylphenyl units to the neighboring positions to obtain 1.37.76 Scheme 1.15. Five-member ring cyclization and aryl rearrangement under Scholl reaction conditions. Taking advantage of the rearrangement, the same authors designed a suitably chlorinated analog 1.38, as shown in Scheme 1.16, to obtain the known rearranged structure 1.39 and perform a palladium-catalyzed cyclization providing the π- extended buckybowl structure 1.40.77 76 J. Liu, A. Narita, S. Osella, W. Zhang, D. Schollmeyer, D. Beljonne, X. Feng, K. Müllen, J. Am. Chem. Soc. 2016, 138, 2602. 77 J. Liu, S. Osella, J. Ma, R. Berger, D. Beljonne, D. Schollmeyer, X. Feng, K. Müllen, J. Am. Chem. Soc. 2016, 138, 8364. Chapter 1 23 Scheme 1.16. Taking advantage of the rearrangement and cyclization to introduce five- membered rings leading to a buckybowl structure. Rearrangements and cyclization can also provide six-membered rings. Müllen et al. evaluated two different polyphenylene isomeric precursors 1.41 and 1.42 (Scheme 1.17) under oxidation conditions. While 1.41 provided the expected planar nanographene 1.43, the polyphenylene 1.42 entailed a 1,2-shift of the central meta- substituted benzene ring to provide also nanographene 1.43 quantitatively.78 Scheme 1.17. Expected graphitization of polyphenylene 1.41, and rearrangement of 1.42 to provide the same nanographene 1.43. Furthermore, Müllen and coworkers reported the cyclodehydrogenation of cyclo- para-phenylene 1.44 yielding to the graphitized cyclo-para-phenylene 1.45 involving a 1,2-phenyl shift induced by the strain in the macrocycle (Scheme 1.18).79 78 M. Müller, V. S. Iyer, C. Kübel, V. Enkelmann, K. Müllen, Angew. Chem. Int. Ed. 1997, 36, 1607. 79 F. E. Golling, M. Quernheim, M. Wagner, T. Nishiuchi, K. Müllen, Angew. Chem. Int. Ed. 2014, 53, 1525. Chapter 1 24 Scheme 1.18. Strain-induced 1,2-phenyl shift. Moreover, the electronic nature of the substitution of the polyarene precursor has an influence on unexpected reactions during the Scholl oxidation. Alkoxy-substituted hexaarylbenzene 1.46, reported by Müllen et al., under Scholl reaction conditions leads to the formation of two products, as shown in Scheme 1.19. The spirocompound 1.47 in which the graphitization is interrupted, and the graphitized hexabenzocoronene 1.48 whose formation results from a rearrangement. The methoxy substituents starting in pseudo-para positions rearrange to pseudo-meta positions in the graphitized structure 1.48.80 Scheme 1.19. The rearrangement induced by the electronic nature of the susbtituents in the polyarene 1.46 yields the spirocompound 1.47 and the unexpected hexabenzocoronene derivative 1.48. 80 X. Dou, X. Yang, G. J. Bodwell, M. Wagner, V. Enkelmann, K. Müllen, Org. Lett. 2007, 9, 2485. Chapter 1 25 In addition, rearrangements can lead to ring contractions. The tetrabenzocyclooctatetraene core of polyphenylene 1.49, shown in Scheme 1.20, under Lewis acid conditions yielded the planar hexagon-constituted nanographene 1.50.81 Scheme 1.20. Ring-contraction rearrangement leading to planar six-membered rings nanographene. Rearrangement followed by seven-membered rings formation was reported by Durola and coworkers (Scheme 1.21). The sterically hindered triarylterphenyl precursor 1.51 in presence of FeCl3 led to the seven-membered containing compound 1.52 resulting from a 1,2-aryl shift.82 Scheme 1.21. Cyclization followed by rearrangement and seven-membered ring formation. A similar observation was reported by Narita et al. (Scheme 1.22). The treatment of the polyarene 1.53 with DDQ/TfOH yielded the unexpected negatively curved nanographene 1.54 containing two seven-membered rings due to a 1,2-naphthyl- phenyl migration. Theoretical calculations supported the evolution through the rearrangement pathway over the formation of the helicene containing graphitized nanographene 1.55.83 81 M. Müller, V. S. Iyer, C. Kübel, V. Enkelmann, K. Müllen, Angew. Chem. Int. Ed. 1997, 36, 1607. 82 A. Pradhan, P. Dechambenoit, H. Bock, F. Durola, J. Org. Chem. 2013, 78, 2266. 83 Z. Qiu, S. Asako, Y. Hu, C.-W. Ju, T. Liu, L. Rondin, D. Schollmeyer, J.-S. Lauret, K. Müllen, A. Narita, J. Am. Chem. Soc. 2020, 142, 14814. Chapter 1 26 Scheme 1.22. Rearrangement by 1,2-naphtyl-phenyl migration and subsequent seven- membered rings cyclization. A few examples of pentagon-heptagon containing nanographenes resulting from rearrangements have been recently reported.84 Feng and coworkers reported the formation of two azulene embedded nanographenes, 1.56 and 1.57 (Scheme 1.23). The formation of the unexpected nanographenes results from the bond rotation in the polyarene precursor 1.58 and a 1,2-aryl migration with subsequent HI elimination.85 Scheme 1.23. 1,2-Aryl migration rearrangement and subsequent HI elimination leading to two azulene-containing nanographenes from different rotamers of the polyarene precursor. The simultaneous ring expansion and contraction leading to azulene-embedded nanographenes has also been reported by Chi et al. Naphthalene-containing polyarene 1.59 in presence of DDQ/TfOH leads to the formation of the two fused embedded azulenes nanographene 1.60 (Scheme 1.24).86 84 J. Ma, Y. Fu, E. Dmitrieva, F. Liu, H. Komber, F. Hennersdorf, A. A. Popov, J. J. Weigand, J. Liu, X. Feng, Angew. Chem. Int. Ed. 2020, 59, 5637. 85 J. Ma, Y. Fu, E. Dmitrieva, F. Liu, H. Komber, F. Hennersdorf, A. A. Popov, J. J. Weigand, J. Liu, X. Feng, Angew. Chem. Int. Ed. 2020, 132, 5686. 86 Y. Han, Z. Xue, G. Li, Y. Gu, Y. Ni, S. Dong, C. Chi, Angew. Chem. Int. Ed. 2020, 59, 9026. Chapter 1 27 Scheme 1.24. Two embedded fused azulenes nanographene. 1.1.3.2 Functionalization During the Scholl Reaction In addition to unexpected rearrangements and cyclizations, functionalization of the products can occur during the Scholl reaction. As observed in our research group (Scheme 1.25), modifying the reaction conditions could entail modifications in the resulting compound. In 2018, the selective formation of corannulene-containing nanographenes 1.61, 1.62 and 1.63 was described.87 When the graphitization reaction is carried out in the presence of FeCl3 under mild conditions (–50 ºC), the [6]helicene embedded nanographene 1.61 is obtained. However, if the reaction is performed at higher temperature (80 ºC), another ring is formed, giving the chlorinated nanographene 1.62 with a seven-membered ring. Additionally, carrying out the reaction with DDQ and triflic acid yields nanographene 1.63, which also contains the seven-membered ring. 87 J. M. Fernández-García, P. J. Evans, S. Medina Rivero, I. Fernández, D. García-Fresnadillo, J. Perles, J. Casado, N. Martín, J. Am. Chem. Soc. 2018, 140, 17188. Chapter 1 28 Scheme 1.25. Selective synthesis of helical, curved chlorinated and curved nanographenes under different Scholl reaction conditions. The possibility of chlorinating electron-rich aromatic compounds in presence of FeCl3 was firstly demonstrated by Niementowski in 1919.88 When performing the Scholl reaction, this chlorination is usually undesired and often unexpected, as it may occur in all-carbon and hydrogen structures. However, different studies have proven the efficiency of the reaction when it occurs simultaneously and selectively on the substrate.89 For instance, Stuparu and coworkers have recently reported the mechanochemical synthesis of corannulene-containing nanographene 1.64 (Scheme 1.26). In a solvent-free assembly of polyarene 1.65 with FeCl3, the selective chlorination and graphitization occurs with very high yield, 90%.90 88 L. T. Bratz and S. von Niementowski, Ber. Dtsch. Chem. Ges. (A and B Ser.), 1919, 52, 189. 89 D. M. Baier, S. Gratz, B. Farhadi Jahromi, S. Hellmann, K. Bergheim, W. Pickhardt, R. Schmid, L. Borchardt, RSC Adv. 2021, 11, 38026. 90 J. Stanojkovic, R. William, Z. Zhang, I. Fernández, J. Zhou, R. D. Webster, M. C. Stuparu. Nat. Commun. 2023, 14, 803. Chapter 1 29 Scheme 1.26. Selective chlorination during mechanochemical Scholl reaction with FeCl3. Some authors suggest the use of MoCl5 to perform the Scholl reaction avoiding chlorination. However, Itami et al. reported the selective tetrachlorination during the graphitization of polyarene 1.66 to give nanographene 1.67 in presence of MoCl5 (Scheme 1.27).91 Scheme 1.27. Selective chlorination during the Scholl reaction with MoCl5. Furthermore, Mastalerz and coworkers reported a double regioselective triflyoxylation and cyclization of pentacene-based structure 1.68 while performing the Scholl reaction with DDQ and triflic acid to obtain structure 1.69 (Scheme 1.28).92 Performing the reaction in presence of FeCl3 under mild conditions, lead to the non- functionalized compound 1.70. Scheme 1.28. Selective triflyoxylation during Scholl reaction with DDQ and triflic acid. Chlorination is not observed when the reaction is performed with FeCl3. 91 T. Fujikawa, N. Mitoma, A. Wakamiya, A. Saeki, Y. Segawa, K. Itami, Org. Biomol. Chem. 2017, 15, 4697. 92 X. Yang, M. Hoffmann, F. Rominger, T. Kirschbaum, A. Dreuw, M. Mastalerz, Angew. Chem. Int. Ed. 2019, 58, 10650. Chapter 1 30 However, the modification of the starting pentacene-based compound functionalized with naphthyl groups 1.71 shows a significant impact on the products obtained.93 When the oxidation reaction is assembled with FeCl3 at reflux of dichloromethane/nitromethane (Scheme 1.29), the selectively chlorinated derivative 1.72 is obtained. In addition, if the reaction is performed with more equivalents of DDQ (8 or 10 equiv) and triflic acid, three different products are obtained. In all cases two seven-membered rings are formed. Nevertheless, the products can be symmetrically substituted with two triflates 1.73, asymmetrically substituted with a triflate and a dichlorovinylidene group 1.74 or symmetrically substituted with two dichlorovinylidene groups 1.75. Scheme 1.29. Selective chlorination in presence of FeCl3. Functionalization with triflate and dichlorovinylene depending on the reaction conditions using DDQ and triflic acid. To conclude, predicting the structure of a nanographene after the Scholl oxidation can be challenging regarding the possible rearrangements and functionalization that may take place during the reaction. The possibilities to modulate the starting structure, the reagents and reaction conditions provide several alternatives to control or conduct the reaction towards different products. Furthermore, the knowledge acquired during the last decade in the synthesis of molecular nanographenes is paving the way for understanding the Scholl cyclodehydrogenation. 93 X. Yang, F. Rominger, M. Mastalerz, Angew. Chem. Int. Ed. 2019, 58, 17577. Chapter 1 31 1.2 Objectives The initial objective of this chapter was the synthesis of a helically arranged nanographene based on two graphitized units covalently connected through a 9,10- substituted anthracene core. However, the methodology used, based on the Diels- Alder reaction to obtain the polyarene, and subsequent Scholl reaction for graphitization, led to the unexpected formation of spironanographenes. Once the formation of the spirocycles was understood and in pursuit of the initial objective, the starting polyarene was modified by substituting the anthracene core with eight fluorine atoms. The electronic effects of these substituents controlled the reaction and the desired helically arranged nanographene was formed. Once the synthesis and characterization were performed, the determination of the optoelectronic properties was carried out by means of absorption, emission, and electrochemical measurements. 32 Chapter 1 33 1.3 Results and Discussion 1.3.1 Unexpected Synthesis of Spironanographenes The initial hypothesis of this work was the synthesis of nanographene 1 consisting of three subunits, two dibenzo[fg,ij]phenanthro-[9,10,1,2,3-pqrst]pentaphene (DBPP) connected by a 9,10-substituted anthracene. The synthetic approach (Scheme 1.30) starts with a double Sonogashira cross-coupling reaction between 9,10- dibromoanthracene (2) and two equivalents of 4-(tert-butyl)phenylacetylene (3), leading to 9,10-bis[(4-(tert-butyl)phenyl)ethynyl]anthracene (4). Subsequently, dialkyne 4 undergoes a two-fold Diels-Alder cycloaddition reaction with two equivalents of cyclopentadienone 5 as the diene, affording polyarene 6a after a CO extrusion. The final step, the graphitization of 6a, was expected to afford the helically arranged nanographene 1. However, the Scholl cyclodehydrogenation led to the formation of two spirocycles connecting the polyarene units to the central anthracene core. The fine control over the Scholl reaction conditions yielded selectively to spirocompounds 7 and 8, and spironanographenes 9 and 10. Modifications in the oxidant, the temperature, and the reaction time allowed to control the number of bonds formed and, therefore, the graphitization degree of the final compound. Chapter 1 34 Scheme 1.30. Synthetic scheme for the synthesis of polyphenylene 6a and further Scholl reaction leading to spirocompounds 7 and 8 (3 new C−C bonds), and spironanographenes 9 (7 new C−C bonds, dashed lines in compound indicates that one of these bonds is not formed) and 10 (8 new C−C bonds). To the best of our knowledge, very few examples of spironanographenes can be found in the literature.94,95 Frequently, the formation of spirocycles during the Scholl reaction 94 S. Liu, D. Xia, M. Baumgarten, ChemPlusChem 2021, 86, 36. 95 X. Zhang, Z. Xu, W. Si, K. Oniwa, M. Bao, Y. Yamamoto, T. Jin, Nat. Commun. 2017, 8, 15073. Chapter 1 35 results from rearrangements or the formation of stable carbocations and entails the interruption of the graphitization process.80,96 However, the graphitization process of the polyarene 6a is not compromised by the formation of the spirocycles with the central anthracene. The use of FeCl3 as oxidant at 85 ºC for 4 hours afforded an isomeric mixture of spirocompounds syn 7 and anti 7. During the evaluation of the reaction under these relatively harsh conditions, complex mixtures of chlorinated compounds were observed by mass spectrometry. This undesired reactivity was prevented by adding a silver salt (AgPF6) that precipitates the chlorides generated during the reaction as silver chloride. Spirocompounds 7 were characterized by the common techniques, NMR, FT-IR and mass spectrometry. MALDI-ToF experiments showed an exact mass of 1649.0632, which indicated the loss of two hydrogen atoms compared to polyarene 6a (HRMS found, 1651.0783). The most relevant NMR information was revealed by 13C-NMR, an apparently unshielded aliphatic signal at 63.8 ppm corresponding to a quaternary carbon (Appendix 5, spectra). Considering the number of signals, the obtained structure might be symmetric, as only signals corresponding to the half of the molecule were observed. The exact determination of the structure was complex by NMR experiments; therefore, it was necessary to resort to single-crystal X-ray diffraction (Appendix 1, crystallographic data). The crystals obtained showed different crystallization shapes allowing their manual separation and independent diffraction. Needle-shaped crystals corresponded to a polyarene-like structure in which two five- membered rings formed spirocycles with the central 9,10-anthracene symmetrically, forcing the orthogonal arrangement of the five-membered rings with the central anthracene. As shown in Figure 1.2a, both 4-tert-butylphenyl rings (in blue) forming the spirocycles (in orange) are on the same face of the anthracene, syn 7. Moreover, ribbon-shaped crystals were resolved and revealed an anti 7 isomer (Figure 1.2b) in which the 4-tert-butylphenyl rings forming the spirocycles are on opposite faces of the central anthracene. 96 J. L. Ormsby, T. D. Black, C. L. Hilton, Bharat, B. T. King, Tetrahedron 2008, 64, 11370. Chapter 1 36 Figure 1.2. Molecular structure of the compounds (a) syn 7 (b) anti 7 determined by single crystal X-ray diffraction, hydrogen atoms have been omitted for clarity. The use of FeCl3 as oxidant was also evaluated at low temperatures. At –30 ºC for 30 minutes, the reaction yielded a partially graphitized isomeric mixture of spirocompound 8. MALDI-ToF experiments showed only one peak with an exact mass of 1643.0344, which corresponds to the loss of eight hydrogen atoms and the subsequent formation of five new bonds. The new structures resulted highly asymmetric and the characterization by NMR was challenging (see Appendix 5). The structure corresponding to syn 8 was resolved by single crystal X-ray diffraction and showed the formation of the previously observed spirocycles and three new C−C bonds on one side of the polyarene, Figure 1.3. The bonds formed are those that do not involve the 4-tert-butylphenyl group forming the spirocycle, the formation of a [5]helicene with an embedded five-membered ring opens the distance between the graphitizable rings, thus hindering the bonding. Figure 1.3. Molecular structure of compound syn 8 solved by single crystal X-ray diffraction, hydrogen atoms have been omitted for clarity. Assembling the reaction with DDQ as oxidant and triflic acid at –60 ºC for 40 minutes, yielded a new apparently graphitized compound 9. Mass spectrometry showed an exact mass of 1634.9775, corresponding to the loss of sixteen hydrogen atoms and the consequent formation of nine bonds, the two expected spirocycles and seven of eight (a) (b) Chapter 1 37 C−C bonds in the polyarene moieties. By analyzing the 1H-NMR spectrum in depth (see Appendix 5), several singlets at lower field indicated the graphitization of the isomers. Finally, the use of DDQ as oxidant was evaluated at higher temperatures. At 40 ºC for 2 hours, the reaction was directed towards the formation of the fully graphitized spironanographene 10 as a single isomer. MALDI-ToF experiments confirmed the formation of ten bonds, (exact mass, 1632.9375) and NMR experiments showed a highly symmetric graphitized structure. 1.3.2 Characterization of Spironanographene 10 The 1D and 2D NMR experiments used for the complete characterization of the new spironanographene are described below. Considering the symmetry of the structure only one half of the molecule is shown, hydrogens have been removed for clarity. Signals will be associated to a coloured dot located in the structure with the goal of assigning the signals to the corresponding hydrogen and carbon atoms to which is bonded. The 1D 1H-NMR spectrum of spironanographene 10 shows the expected signals (Figure 1.4): at low field eight singlets assignable to the aromatic protons on the graphitized side (green dots), four multiplets corresponding to the anthracene moiety (blue dots), a spin system composed by a doublet (yellow dot), a doublet of doublets (red dot), and a long distance doublet with small 4JH–H (orange dot), corresponding to the ring opposite to the spirocycle in the graphitized unit, and at high field five singlets (purple dots) corresponding to the tert-butyl groups. 5 tBu Aliphatic singlets 8 Aromatic singlets 4 Multiplets 1 Doublet 1 Doublet of doublets 1 Doublet (small 4JH-H) (a) Chapter 1 38 Figure 1.4. (a) Structure and the expected signals for 1H-NMR indicating the number and multiplicity. (b) Low field 1H-NMR signals. (c) High field 1H-NMR signals corresponding to nanographene 10. (d) Assignations considering the 1H-NMR spectrum. Considering the multiplicity, the coupling constants and the integration, the following protons can be assigned in the structure, Figure 1.4d. The proton at 8.23 ppm (doublet, orange dot) is coupled to the proton at 5.80 ppm (double doublet, red dot) with a 4JH– H of 2.0 Hz. In turn, the proton at 5.80 ppm (double doublet, red dot) shows a coupling with the proton at 7.99 ppm (doublet, yellow dot) with a 3JH–H of 8.6 Hz. In Figure 1.5 the 13C and DEPT NMR spectra are shown. The large number of carbons present in the structure usually makes the analysis of these experiments complex. Thirty-six of the expected forty-two sp2 carbons are observed, as well as the ten expected aliphatic carbons corresponding to the tert-butyl groups (five quaternary carbons and five assigned to the −CH3). Additionally, the key signal for the assignation of spironanographene 10 is observed at 69.9 ppm (circled in light green). This unshielded aliphatic signal corresponds to the quaternary sp3 spiranic carbon. 1H NMR (500 MHz, CDCl3) J=8.6 Hz J=2.0 Hz J=2.0 Hz J=8.6 Hz (b) (c) 5.80 8.23 7.99 (d) Chapter 1 39 Figure 1.5 13C and DEPT NMR spectra corresponding to spironanographene 10. Assignation of the spiranic carbon. By means of the HMQC (Heteronuclear Multiple Quantum Correlation) experiment, shown in Figure 1.6, the carbons interacting by a 1JC–H coupling constant with hydrogens could be assigned. At this point only the carbons interacting with the previously assigned hydrogens are indicated and the others will be added later as the structure is resolved. Therefore, the following carbons are assigned considering the crossings: 8.23 ppm /118.3 ppm (orange dot), 7.99 ppm / 131.1 ppm (yellow dot), 5.80 ppm /122.8 (red dot). 5.80 8.23 7.99 69.9 Chapter 1 40 Figure 1.6 HMQC 13C-1H NMR experiment corresponding to spironanographene 10. Assignation of carbons directly bonded to hydrogens. The HMBC (Heteronuclear Multiple Bond Correlation) experiment shows some relevant crossings 13C–1H at 3JC–H. The crossings between the spiranic carbon at 69.9 ppm and the following aromatic hydrogens are key to determine the structure, Figure 1.7. The singlet at 7.78 ppm interacting at 3JC–H with the spiranic carbon corresponds to the hydrogen in the phenyl ring involved in the spirocycle (water blue dot, the carbon can also be assigned considering the HMQC in Figure 1.6, 7.78-122.1), the multiplet at 7.00 ppm (light-blue dot) corresponds to one of the hydrogens at three bonds distance from the spiranic carbon in the anthracene, and the multiplet at 6.66 ppm (light-blue dot) corresponds to the other hydrogen at three-bonds distance in the anthracene from the spiranic carbon. 5.80 122.8 8.23 118.3 7.99 131.1 69.9 Chapter 1 41 Figure 1.7 HMBC 13C-1H NMR experiment corresponding to spironanographene 10. Assignation of hydrogens interacting at three bonds distance (3JC–H) with the spiranic carbon. The crossings observed in HMBC between the quaternary aliphatic carbons, and the aromatic hydrogens allow further assignation of the tert-butyl groups on each phenyl ring and one more aromatic hydrogen, as well as the confirmation of the ones located so far, Figure 1.8. The aromatic hydrogens at 8.23 (orange dot) and 5.80 (red dot) ppm show interaction with the quaternary carbon at 34.0 ppm (orange cross). The hydrogens at 8.65 (light green dot) and 7.78 (water blue dot) ppm interact with the quaternary carbon at 36.3 ppm (blue cross). 5.80 122.8 8.23 118.3 7.99 131.1 69.9 7.00 6.66 7.78 122.1 Chapter 1 42 Figure 1.8 HMBC 13C-1H NMR experiment corresponding to spironanographene 10. Assignation of aliphatic carbons interacting at three bonds distance (3JC–H) with hydrogens. The assignation is complemented by homonuclear scalar coupling experiment 1H-1H COSY (COrrelated SpectroscopY) that shows the interaction between coupled protons, given the rigidity of the structure the interaction between protons at 4JH–H are clearly observed, Figure 1.9. The spin system formed between the protons at 8.23, 5.80 and 7.99 ppm is confirmed, as well as the system formed by the protons at 7.78 and 8.65 ppm (COSY interaction highlighted in blue). Furthermore, it is observed a new coupling for further assignation between the anthracene protons at 7.00 and 6.66 ppm with the multiplet at 6.61 ppm (dark blue dot, carbon assigned considering the previously showed HMQC in Figure 1.6). One more interaction between the protons at 9.23 and 8.82 ppm is observed but not yet assignable. 5.80 122.8 8.23 118.3 7.99 131.1 69.9 7.00 6.66 7.78 122.1 34.03 36.3 8.65 Chapter 1 43 Figure 1.9 1H−1H COSY NMR experiment corresponding to spironanographene 10. Assignation of aromatic hydrogens showing scalar coupling. In these rigid systems in which the protons are located at the periphery of the structure, dipole (through space) couplings are usually observed. Therefore, NOESY (Nuclear Overhauser Effect Spectroscopy) experiments were performed, Figure 1.10. Analyzing the interactions between the aromatic protons, those sharing the bay-region of the graphitized area are close enough to show interaction (indicated with yellow arrows). Thus, the proton at 8.65 ppm can be assigned by the interaction with the proton at 9.07 ppm. Same interaction occurred between protons at 8.82 and 8.23 ppm, once located the proton at 8.82 ppm, the one at 9.23 ppm is confirmed by the previously observed COSY interaction (Figure 1.9). 5.80 122.8 8.23 118.3 7.99 131.1 69.91 7.00 6.66 7.78 122.1 34.0 36.3 8.65 6.61 128.2 6.61 128.2 9.23 8.82 Chapter 1 44 Figure 1.10 NOESY NMR experiment corresponding to spironanographene 10. Assignation of aromatic hydrogens showing dipolar coupling in the bay regions of the graphitized area. In addition to interactions between aromatic protons, the NOESY experiment also shows interactions between protons in the tert-butyl groups (highlighted in pink) and the aromatic protons of the same ring, as shown in Figure 1.11. The interactions between the protons at 9.23 and 8.82 ppm with the aliphatic singlet at 1.77 ppm locates the tert-butyl in the ring. The same interactions are observed between the protons at 8.65 and 7.78 ppm with the singlet at 1.50 ppm, and between the protons at 8.23 and 5.80 ppm with the aliphatic signal at 0.72 ppm. 5.80 122.8 8.23 118.3 7.99 131.1 69.91 7.00 6.66 7.78 122.1 34.0 36.38.65 6.61 128.2 6.61 128.2 9.23 8.82 9.07 NOESY NOESY Chapter 1 45 Figure 1.11 NOESY NMR experiment corresponding to spironanographene 10. Assignation of tert-butyl groups showing dipolar coupling with the aromatic hydrogens in the same ring. So far, only interactions between atoms in the same graphitized unit of the structure have been described. This allows the confirmation of the structure; however, it remains to identify whether the syn-isomer or the anti-isomer has been obtained. The key in the identification of the isomer is observed in the dipolar interaction between the aromatic proton at 8.82 ppm and the tert-butyl group at 0.72 ppm, as reflected by the corresponding crossover in the NOESY spectrum, Figure 1.11. The dipolar coupling between the tert-butyl (highlighted in pink) and the aromatic hydrogen (highlighter in green) can only result from the interaction between the two graphitized units. Considering the disposition of both possible syn/anti isomers and the location of the protons (Figure 1.12), the through-space coupling can only occur 5.80 122.8 8.23 118.3 7.99 131.1 69.91 7.00 6.66 122.1 7.78 34.0 36.38.65 6.61 128.2 6.61 128.2 9.23 8.82 9.07 NOESY0.72 1.77 1.50 Chapter 1 46 in syn isomer. Thus, the structure obtained from the Scholl reaction is the spironanographene syn 10. Figure 1.12 Location of the through-space coupled protons (highlighted in green and pink) in the corresponding syn and anti-isomers of spironanographene 10. In addition to the complete characterization of spironanographene syn 10 by NMR, it was possible to resolve the structure by single-crystal X-ray diffraction. As shown in Figure 1.13, it is confirmed the formation of the syn isomer, in which the graphitized units are arranged orthogonally with respect to the central anthracene. Due to steric hindrance between the two units a helical organization is formed across the spiranic rings. Figure 1.13 Molecular structure of the compounds syn 10 resolved by single crystal X-ray diffraction, hydrogen atoms have been omitted for clarity. Chapter 1 47 1.3.3 Mechanism of the Scholl reaction To further understand the unexpected formation of spironanographene syn 10 against the formation of compound 1, DFT calculations at the dispersion-corrected PCM(CH2Cl2)-B3LYP-D3/def2-SVP level were carried out. Figure 1.14 Computed reaction profile for the Scholl and spiro-cyclization reactions involving anthracene derivative 6aM as computational model. Free energy values (ΔG, 298 K) are given in kcal/mol. Negative numbers in blue indicate the computed NICS(1)zz values (in ppm). All data have been computed at the PCM(CH2Cl2)-B3LYP-D3/def2-SVP level. In Figure 1.14 is represented the computed reaction profile considering the formation of cation intermediates, the formation of radical-cation intermediates leads to the same conclusions (see Appendix 1, theoretical calculations), according to the computed barrier and reaction energies, it can be concluded that the arenium cation pathway is favored over the radical cation mechanism. Starting from the model polyarene 6aM, two possible carbocations can be formed under Scholl reaction conditions (FeCl3 or DDQ/TfOH). The expected protonation at one of the aryl groups in the periphery of the polyarene would provide the intermediate INT1’. However, the formation of the Chapter 1 48 intermediate INT1 if the protonation occurs directly at the anthracene core is strongly preferred (ΔΔG= 13.0 kcal/mol) over the expected Scholl intermediate INT1’. Additionally, the subsequent Scholl-like cyclization from INT1’, via transition state TS1’, is severely hampered in comparison to the alternative cyclization of INT1 toward the spirocycle of INT2 (both kinetically ΔΔG≠= 14.6 kcal/mol and thermodynamically ΔΔGR= 19.2 kcal/mol). Therefore, leaving evidence of the preferred reaction pathway leading to spirocycles over the expected Scholl reaction. Which is consistent with the formation of partially graphitized spirocompounds 7, 8 and 9 and the totally graphitized spironanographene 10. The reasons for the enhanced stability of the initial intermediate INT1 versus the graphitization intermediate INT1' is strongly related to the aromaticity of the involved species. Nuclear Independent Chemical Shift (NICS) calculations indicate that the aromaticity of the peripheral aryl group in the initial model 6aM (NICS(1)zz= –31.3 ppm) is dramatically reduced in the protonated intermediate INT1’ (NICS(1)zz= –13.4 ppm). Whereas, the aromaticity of the lateral ring of the anthracene core (NICS(1)zz= –22.4 ppm) increases after the formation of a trityl cation (in orange) in the central ring (NICS(1)zz= –26.9 ppm). The formation of a new Clar-sextet in the system enhances the thermodynamic stability, being the driving force of the transformation. Therefore, the highly stable trityl cation in INT1, LUMO mainly located in the carbon bearing the positive charge, evolves to INT2 after a SEAr reaction with one of the closest peripheral rings in the polyarene. Seeking a structure similar to the target nanographene 1, we considered the possibility of decorating the central anthracene with electron withdrawing atoms. NICS(1)zz calculations indicated a higher aromaticity (–24.9 ppm) for the hypothetical octafluorinated anthracene derivative 6bM than for the cationic specie INT1b (–20.7 ppm). Thus, the evolution from 6bM to INT1b involves an aromaticity loss (ΔNICS(1)zz = –4.2 ppm), against the aromaticity gain previously observed for 6aM, which may prevent the presence of the spirocyclization pathway. 1.3.4 Electronic Effects Controlling the Scholl Reaction In section 1.3.1 the unexpected obtaining of spironanographenes with different degrees of graphitization has been described. The mechanistic theoretical calculation showed the formation of very stable trityl cation implying an aromaticity gain, which makes polyarene 6a evolve towards the formation of spirocycles. However, it might be expected that the formation of the trityl carbocation in the anthracene core could be avoided by introducing electron-withdrawing atoms, therefore leading to a derivative of the goal nanographene 1. Motivated by the theoretical evidence, the synthesis of polyarene 6b, in which the pentaarylbenzene units are connected by an octafluorinated anthracene, was designed. As shown in Scheme 1.31, the synthesis starts with a double Sonogashira coupling between 9,10-dichloro-1,2,3,4,5,6,7,8-octafluoroanthracene (11) and two equivalents Chapter 1 49 of 4-(tert-butyl)phenylacetylene (3), leading to 9,10-bis[(4-(tert-butyl)phenyl) ethynyl]-1,2,3,4,5,6,7,8-octafluoroanthracene (12). The obtained dialkyne 12 undergoes a Diels-Alder cycloaddition followed by CO extrusion with cyclopentadienone 5, affording the polyarene 6b. The final Scholl oxidation of the fluorinated polyarene 6b using DDQ as oxidant in presence of triflic acid yields the expected nanographene 13. In agreement with the theoretical calculations, the electron-withdrawing effects of the fluorine atoms decorating the anthracene core control the Scholl reaction by preventing the formation of spirocycles. Thus, yielding nanographene 13, in which the graphitized units and the fluorinated anthracene are helically arranged. Scheme 1.31. Synthetic scheme for the preparation of polyphenylene 6b and further Scholl reaction leading to helically arranged nanographene 13. Chapter 1 50 Structure 13 was fully characterized by NMR (1H, 13C and 19F), FT-IR and high- resolution mass spectrometry. As shown in Figure 1.15, the great symmetry of the structure, three C2 axes, is reflected in the NMR spectra. The spectra show one quarter of the molecule; particularly the signals in 1H-NMR are one singlet at 9.22 ppm (red dot), four doublets with 4JH-H= 1.7 Hz at 9.17, 8.99 and 8.70 ppm (orange dots), a doublet at 7.55 ppm with a 3JH-H= 8.9 Hz (blue dot), a doublet of doublets at 6.99 ppm (3JH-H= 8.9 Hz and 4JH-H= 1.7 Hz, green dot) and three aliphatic singlets corresponding to the tert-butyl groups at 1.83, 1.78 and 1.20 ppm, with relative intensities 1:2:2. Additionally, the 19F-NMR spectra also reveal the symmetry, only two doublets (3JF- F= 14.4 Hz) are observed. The structure was modelled and showed the expected helically arrangement of the moieties, as shown in Figure 1.15. Figure 1.15. Modelization of helically arranged nanographene 13 and its 1H NMR (500 MHz, CDCl3) and 19F NMR (471 MHz, CDCl3) spectra. 1.3.5 Electrochemical and Photophysical Properties The electrochemical properties of spironanographene syn 10 and helically arranged nanographene 13 were evaluated by cyclic voltammetry (data summarized in Table 1.1). The measurements were performed in a 1 M solution of tetrabutylammonium hexafluorophosphate as supporting electrolyte in toluene/acetonitrile 1:1, using Ag/AgNO3 as reference electrode at room temperature. Chapter 1 51 In Figure 1.16 are represented the voltamperograms corresponding to spironanographene syn 10 and hexa-tert-butyl-hexa-peri-hexabenzocoronene (tBu- HBC) vs Fc/Fc+. As observed, spironanographene syn 10 is a poorer electron acceptor than tBu-HBC considering the first quasireversible reduction potentials, E1 1/2red= –2.56 and E1 1/2red= –2.27 V, respectively. Moreover, considering the first oxidation potentials, E1 1/2ox= 0.85 V for syn 10 and E1 1/2ox= 0.67 V for tBu-HBC, it can be also concluded that the spironanographene is a weaker electron donor than tBu-HBC. Figure 1.16. Cyclic voltammograms corresponding to spironanographene syn 10 (top) and hexa-tert-butyl-hexa-peri-hexabenzocoronene (tBu-HBC, bottom) vs Fc/Fc+ in a 1 M solution of tetrabutylammonium hexafluorophosphate in toluene/acetonitrile 1:1. The higher electronic band gap of the spironanographene (syn 10 EGAP= 3.13 eV vs tBu-HBC 10 EGAP= 2.65 eV) can be explained in terms of the π-extension differences. The formation of the spiranic rings entails the disruption of the conjugation, whereas in HBC the six rings attached to the central benzene are graphitized and therefore conjugated, in spironanographene 10 the conjugation occurs only between five of the rings attached to the central benzene. Although the spironanographene presents two graphitized layers, the electronic communication is hampered by the sp3 carbon atoms forming the spirocycles. Chapter 1 52 The electrochemical properties of the fluorinated helically arranged nanographene 13 were also evaluated by cyclic voltammetry. As shown in Figure 1.17, the electron acceptor character is stronger than that of tBu-HBC, the first quasireversible band appears at E1 1/2red= –1.69 V. This significant shift to less negative values respect to HBC (E1 1/2red= –2.27 V), is due to the important electron-withdrawing character of the perfluorinated anthracene connecting the graphitized units. Further reduction waves were observed at E2 1/2red= –2.40, E3 1/2red= –2.47 and E4 1/2red= –2.69 V, indicating that nanographene 13 can accept more electrons under these conditions. Regarding the donor character, nanographene 13 presents one quasireversible oxidation wave at E1 1/2ox= 0.87 V, showing a slightly poorer donor character than tBu-HBC (0.67 V). This result could be explained by the reduced π-extension of nanographene 13 vs tBu- HBC, and by the nearby bonding of the fluorinated anthracene core, which results in an electronic band gap lower than that of tBu-HBC (13, EGAP= 2.37 eV vs. tBu-HBC, EGAP= 2.65 eV). Figure 1.17. Cyclic voltammograms corresponding to helically arranged nanographene 13 vs Fc/Fc+ in a 1 M solution of tetrabutylammonium hexafluorophosphate in toluene/acetonitrile 1:1. Table 1.1. Redox potentials of tBu-HBC, syn 10 and 13 vs Fc/Fc+. Compound E1 red (V) E2 red (V) E3 red (V) E4 red (V) E1 ox (V) tBu-HBC –2.27 –2.50 –2.79 – 0.67 syn 10 –2.56 – – – 0.85 13 –1.69 –2.40 –2.47 –2.69 0.87 The photophysical properties (data summarized in Table 1.2) of the spironanographene 10 were evaluated by UV-vis absorption and emission experiments. As shown in Figure 1.18 (left), spironanographene syn 10 absorption spectrum presents three well-structured, intense sharp bands (324, 336 and 350 nm) and two weaker bands (276 and 391 nm) in the UV region. In the visible region two Chapter 1 53 weak bands are observed at 414 and 438 nm. The emission spectrum shows three well- structured bands at 439, 456 and 464 nm, and two weaker bands at 484 and 496 nm. Using tBu-HBC as reference, the absorption spectra show similar shape, with three intense sharp bands in the UV region (344, 360 and 390 nm) and three weak bands in the visible region (439, 441 and 443 nm). The emission spectrum shows a broad band and a shoulder at 493 and 553, respectively. The comparison between spironanographene 10 and tBu-HBC show the same tendency previously concluded from the voltammetry measurements. The absorption and emission spectra of nanographene 10 are blue-shifted vs tBu-HBC, which seems reasonable considering the lower π-extension of the spironanographene and the hampered conjugation between the graphitized layers. The calculations of the frontier molecular orbitals (Figure 1.18, right) reinforce these results. Both orbitals HOMO and LUMO are delocalized in both π-extended layers, with no coefficients in the central anthracene core. Furthermore, the TD-DFT calculations indicate that the weak bands in the visible region of the absorption spectrum of 10 (414 and 438 nm) have rather low oscillator strengths (0.06 and 0.09, respectively) because of the vertical transitions’ combinations involving the orbitals from HOMO–4 and LUMO+3. Figure 1.18. Absorption and emission spectra of spironanographene syn 10 (left). Computed frontier orbitals HOMO (top right) and LUMO (bottom right) of syn 10 considering a model with methyl groups instead of tert-butyl groups. The photophysical properties of helically arranged nanographene 13 are represented in Figure 1.19. The absorption spectrum shows three intense sharp bands in the UV region at 310, 337 and 367 nm, and one weaker band at 381 nm. In the visible region three broad less intense bands are observed at 413, 454 and 482 nm. The spectrum shows red-shifted same-shaped bands compared to tBu-HBC (sharp intense bands, Chapter 1 54 344, 360 and 390 nm, and weak bands, 439, 441 and 443 nm). To understand the nature of the bathochromic shift of 13 vs tBu-HBC, TD-DFT calculations were carried out. PCM(hexane)-TD-B3LYP-D3/def2-SVP//B3LYP-D3/def2-SVP calculations on the model 13M revealed that the absorption at 482 nm is assigned to the vertical transition involving a one-electron promotion from HOMO to LUMO. These π- orbitals are mainly located at the fluorinated anthracene core (Figure 1.19, right). However, the band at 454 nm is associated to the transition from the HOMO–1, π- delocalized in the graphitized layers and nearly degenerated with the HOMO, to LUMO. Figure 1.19. Absorption and emission spectra of helically arranged 13 (left). Computed frontier orbitals of fluorinated nanographene 13M considering a model (right). Table 1.2. UV-vis absorption and emission maxima of tBu-HBC, syn 10 and 13 Compound Absorption λabs max (nm) Emission λemi max (nm) tBu-HBC 344, 360, 390, 439, 441, 445 493, 519, 553 Syn 10 324, 336, 350, 376, 391, 414, 438 439, 456, 465, 484, 496 13 310, 337, 367, 381, 413, 454, 482 437, 454, 464, 499, 531, 575 Chapter 1 55 The emission spectra of nanographene 13 shows two weak sharp bands at 437 and 464 nm, and more intense broad bands at 501, 531, 575 nm. The unexpected overlapping between the absorption and emission spectra suggested an apparent dual emission in which the fluorophore fragments in the molecule (fluorinated anthracene and graphitized layers) provide emission signals independently. By comparison with the emission spectra of the precursors, the broad bands at longer wavelengths agree in shape with the emission spectra of 6b, 11 and 12, as shown in Figure 1.20 (left). Figure 1.20. Emission spectra of precursors 11, 12, 6b and nanographene 13 (left). Excitation spectra of nanographene 13 at the emission spectrum maxima wavelengths. To confirm that the signals in the emission spectra resulted from the same molecule and not from impurities, excitation spectra were recorded (Figure 1.20, right). As the spectra obtained did not depend on the selected wavelength, at the expense of an in- depth photophysical study, it could be in principle concluded that the structure shows an apparent dual emission. Further evidence is provided in the discussion of the following Chapter 2.3.6. 56 Chapter 1 57 1.4 Conclusions Controlling the Scholl reaction can be challenging considering the possible rearrangements, unexpected cyclizations or functionalization that may occur under oxidation conditions. In this chapter, the electronic control over the Scholl reaction by modulating the electron-withdrawing character of the central anthracene core connecting two graphitized moieties is described. Anthracene-based polyphenylene 6a led to the unexpected formation of spirocompounds 7, 8, 9 and 10 with different degrees of graphitization depending on the oxidant, temperature, and reaction time, being the first example of spirocycles formation under Scholl reaction conditions without interruption of the graphitization. DFT calculations nicely explained the formation of the spirocycles considering the formation of arenium cation or radical-cation intermediates. The cyclization of the spirocycles precedes the graphitization and the key is the formation of a trityl cation intermediate that involves an aromaticity gain, providing a much favorable pathway than the expected graphitization. In contrast, if the anthracene core is substituted with eight electron-withdrawing fluorine atoms, the Scholl reaction proceeds through graphitization without forming the spirocycles 13. The electronic deficiency in the anthracene hinders the formation of the trityl cation, then the reaction evolves through the expected aryl-aryl coupling. Electrochemical and photophysical properties of the new spironanographene syn 10 and helically arranged nanographene 13, showed the electron donor and acceptor character of the structures. Considering tBu-HBC as reference (E1 1/2ox= 0.67 V and E1 1/2red= –2.27 V), the conjugation interruption by the spiranic sp3 carbon atoms connecting the graphitized units entails the separation between oxidation (E1 1/2ox= 0.85 V) and reduction (E1 1/2red= –2.56V) bands, and the blueshift of the electronic bands, resulting in a bandgap opening for nanographene syn 10. However, helically arranged nanographene 13 presents a lower bandgap considering the first oxidation (E1 1/2ox= 0.87 V) and reduction (E1 1/2red= –1.69 V) potentials, this behavior is confirmed by the red-shifted bands in the absorption and emission spectra. Furthermore, nanographene 13 shows a dual emission with bands apparently steaming from both fluorophores, the octafluoroanthracene and the graphitized moieties. Chapter 2 Chapter 2 61 2 Chapter 2 2.1 Introduction to Chirality in Nanographenes The first experimental evidence of chirality date back to 1808, when Étienne-Louis Malus discovered the polarized light;97 just a few years later, in 1811 Dominique F. J. Arago discovered optical rotation in sheets of quartz crystals.98 In 1817, Baptiste Biot observed optical activity from solutions of different natural products.99 However, the finding of dissymétrie and enantiomerism is acknowledged to Louis Pasteur, who in 1848 manually separated two enantiomorphous crystals of tartaric acid, dissolved them separately and registered their optical activity, finding the same values in absolute magnitude but opposite in direction.100 The following milestone for the understanding of chirality was the determination of the three-dimensional arrangement of the atoms in a molecule. In 1874, Jacobus H. van’t Hoff discovered and defined the asymmetric carbon atom and illustrated two specular images of a tetrahedral carbon, as represented in Figure 2.1a.101 The term of chirality was later stablished by Sir William Thomson, Lord Kelvin, in 1907 (Figure 2.1b)102 “I call any geometrical figure, or group of points, chiral, and say that it has chirality, if its image in a plane mirror, ideally realized, cannot be brought to coincide with itself.” Using the classic example of hands (in Greek χείρ −kheir−, origin of the word chiral) it is shown how the left and the right hands are non-superposable mirror images, Figure 2.1c. Figure 2.1. a) J. H. van’t Hoff’s enantiomers illustration of an asymmetric carbon atom. b) Sir W. Thomson (Lord Kelvin) definition of chirality. c) Hands, non-superposable mirror images. The increasing complexity of the structures, and the difficulties encountered in the nomenclature and identification of chiral molecules led the researchers Robert S. Cahn, Christopher K. Ingold and Vladimir Prelog to join forces and set the priority 97 S. Mauskopf, Trans. Am. Philos. Soc. 1976, 66, 56. 98 S. Mauskopf, Trans. Am. Philos. Soc. 1976, 66, 61. 99 J.-B. Biot, Mém. Acad. R. Sci. Inst. Fr. 1817, 2, 4. 100 L. Pasteur, C. R. Séances Acad. Sci. 1848, 26, 535. 101 J. H. van’t Hoff, Arch. Neerl. 1874, 9, 1. 102 Lord Kelvin, J. Oxford Univ. Junior Sci. Club 1894, 18, 25. (a) (b) (c) Chapter 2 62 rules for the identification of chiral systems.103 In addition, they clarified the definition of chirality stated by Lord Kelvin, establishing that the necessary and sufficient condition for the existence of optical enantiomers is that a mirror plane converts the model into a non-identical one which cannot be superposed on the original by translations and rotations only. Therefore, a model which has no symmetry elements except for rotational axes may be chiral.104 Thus, chirality can be central, axial, and planar, depending on whether the chiral element is an atom, an axis, or a plane, respectively (Figure 2.2). The definitions provided by the IUPAC Gold Book105 (International Union of Pure and Applied Chemistry) for the chiral elements are: • Chirality centre: “An atom holding a set of ligands in a spatial arrangement which is not superposable on its mirror image”. • Chirality axis: “An axis about which a set of ligands is held so that it results in a spatial arrangement which is not superposable on its mirror image”. • Chirality plane: “A planar unit connected to an adjacent part of the structure by a bond which results in restricted torsion so that the plane cannot lie in a symmetry plane”. Figure 2.2. Non-superposable mirror images containing an asymmetric carbon atom (central chirality), a chirality axis (axial chirality, e.g., allenes and atropisomerism of ortho-substituted biphenyls) or a chirality plane (planar chirality, e.g. cyclophanes). Chiral elements are represented in yellow. 103 (a) R.S. Cahn, C.K. Ingold and V. Prelog. J. Chem. Soc. 1951, 612. (b) R.S. Cahn, C.K. Ingold and V. Prelog. Experientia 1956, 12, 81. 104 (a) R.S. Cahn, C.K. Ingold and V. Prelog. Angew. Chem. Inter. Ed. 1966, 5, 385. (b) V. Prelog, G. Helmchen, Angew. Chem. Int. Ed. 1982, 21, 567. 105 IUPAC. Compendium of Chemical Terminology, 2nd ed. (the "Gold Book"). Compiled by A. D. McNaught and A. Wilkinson. Blackwell Scientific Publications, Oxford (1997). *Chirality centre * Chirality axis * Chirality plane Chapter 2 63 2.1.1 From Two-Dimensionality to Chirality Chirality is related to the geometry of the molecule and its organization in the three dimensions of space. However, the nature of the Csp2 atoms that comprise the hexagonal graphene-like networks entails the formation of planar structures implying the presence of a symmetry plane (Figure 2.3a). The required deviation from planarity to generate asymmetry in the search for inherent chirality in nanographenes has been achieved by introducing topological defects following two main approaches. On the one hand, the introduction of non-hexagonal rings surrounded by fused hexagonal rings. Inner rings smaller than six members (e.g. [5]circulene or corannulene) provide bowl-shape structures with positive Gaussian curvature. Inner rings larger than six members lead to saddle-shape structures with negative Gaussian curvature (e.g. [7]circulene, Figure 2.3b).106 On the other hand, the generation of sterically congested structures can induce the torsion of the hexagonal networks if the energy cost of the torsional strain, or deformation, minimizes the steric interactions (e.g. helicenes, twistacenes, and nanobelts, Figure 2.3c).107 Figure 2.3. (a) Csp2 atoms in a hexagonal network provide planar structures with a symmetry plane (σh), in light pink, preventing chirality. (b) *Non-hexagonal rings diverting the structures from planarity, positive Gaussian curvature (bowls) and negative Gaussian curvature (saddles). (c) Strain-induce deviation from planarity. Chiral axes are represented in yellow. *Corannulene and [7]circulene should be substituted to fulfill the symmetry conditions for chirality. Considering the sp2 nature of the carbon atoms, the main chiral elements in molecular nanographenes, to fulfill the symmetry requirements (absence of symmetry planes and inversion centers), are chirality axes or chirality planes (represented in yellow in Figure 2.3). 106 M. Rickhaus, M. Mayor, M. Juríček, Chem. Soc. Rev. 2017, 46, 1643. 107 M. Rickhaus, M. Mayor, M. Juríček, Chem. Soc. Rev. 2016, 45, 1542. Chapter 2 64 2.1.2 Chiral Nanographenes and Isomerization Barriers The isomerization processes of the structural features that induce chirality in nanographenes (Figure 2.3) do not involve the breaking of covalent bonds. High flexibility and molecular dynamics can lead to isomerization. Therefore, additionally to the symmetry requirements, the structure must be stereochemically rigid to be chiral. Understanding the resistance preventing the isomerization (isomerization energy barrier) is key to determine if the chiroptical properties are stable in time, proving the stereochemical rigidity. According to the isomerization energy barriers, molecular nanographenes can be classified in the following dynamic subclasses: flexible, detectable, isolable, or rigid. 2.1.2.1 Flexible Nanographenes (<10 kcal·mol−1) Isomerization energy barriers lower than 5 kcal·mol−1 involve fast conformational equilibrium making enantiomeric species indistinguishable. The presence of seven- membered rings usually means low isomerization barriers given the high flexibility of this structural feature, for example the saddle-to-saddle inversion of [7]circulene is only 0.05 kcal·mol−1. However, these values can be increased by different structural features. The negative curvature of nanographene 2.1 arises from the two seven- membered rings, each one surrounded by six fused rings (four hexagonal and two pentagonal), Figure 2.4. The calculated enantiomerization barrier of 5.9 kcal·mol−1 shows how six fused rings, out of seven possible, tend to increase the value of the isomerization barrier due to the steric hindrance between the hydrogen atoms in the bay region.108 Another topological defect usually related to low isomerization barriers is [4]helicene. The steric hindrance between the hydrogens in the terminal rings is sufficient to divert the structure from planarity, but the rapid interconversion makes the isomers impossible to be detected. Thus, the enantiomers of nanographene 2.2 (Figure 2.4) were only resolved in solid state.109 Figure 2.4. Flexible nanographenes with isomerization barriers lower than 5 kcal·mol−1. 108 K. Oki, M. Takase, S. Mori, A. Shiotari, Y. Sugimoto, K. Ohara, T. Okujima, H. Uno, J. Am. Chem. Soc. 2018, 140, 10430. 109 J. Liu, B.-W. Li, Y.-Z. Tan, A. Giannakopoulos, C. Sánchez-Sánchez, D. Beljonne, P. Ruffieux, R. Fasel, X. L. Feng, K. Müllen, J. Am. Chem. Soc. 2015, 137, 6097. Chapter 2 65 2.1.2.2 Spectroscopically Detectable Chirality (10-20 kcal·mol−1) An isomerization barrier of 20 kcal·mol−1 corresponds to a half-time life of τ1/2= 1000 s at 298 K, which, in most cases, allows the detection of enantiomeric species by spectroscopic methods. The majority of nanographenes with Gaussian curvature are framed in this range. As previously mentioned, the isomerization barriers of [n]circulenes with n−1 fused rings tend to be higher than those of the equivalent [n]circulenes. As shown in Figure 2.5, the energy for the saddle-to-saddle inversion of hexa[7]circulene 2.3 (analog to [7]circulene with only six benzene fused rings) is 17 kcal·mol−1 whereas, for [7]circulene, the saddle-to-saddle barrier is 0.05 kcal·mol−1. Additionally, greater bowl depth, for positively curved nanographenes, or greater saddle depth, for negatively curved nanographenes, increase the energy of the bowl-to-bowl or saddle- to-saddle isomerization barriers. In Figure 2.5a is represented a curved molecular nanographene 2.4 consisting of five hexa[7]circulene units fused to a central corannulene or [5]circulene.110 Thus, the structure presents negative curvature in the periphery and positive curvature in the core. The isomerization barriers of the different topological defects were calculated. The energy value for the saddle-to-saddle inversion of the seven-membered rings (13.6 kcal·mol−1 by NMR and 18.9 kcal·mol−1 by DFT) agrees with the previously described trend of hexa[7]circulenes. In addition, the bowl-to-bowl inversion of the core, with a bowl depth of 0.37 Å, resulted of 1.7 kcal·mol−1, expectedly lower than that of corannulene, 10.4 kcal·mol−1, considering the larger bowl depth, 0.87 Å. Figure 2.5. (a) [n]Circulenes with n-1 fused rings and bowl depth effects over the isomerization barriers of a curved nanographene. (b) Saddle depth effect in negatively curved nanographenes. 110 K. Kawasumi, Q. Zhang, Y. Segawa, L. T. Scott, K. Itami, Nat. Chem. 2013, 5, 739. Chapter 2 66 The effect of the saddle depth is observed in [7]circulene derivatives 2.5 and 2.6, represented in Figure 2.5b. The peripheral π-extension of the negatively curved structure 2.5 involves the generation of two fused [4]helicenes leading to a saddle depth of 2.5 Å and an isomerization barrier of 12.2 kcal·mol−1.111 However, a larger π-extension yielding seven fused [4]helicene in structure 2.6 results in a lower saddle depth, 1.7 Å, and therefore, a lower isomerization barrier, 3.8 kcal·mol−1.112 2.1.2.3 Isolable Chiral Nanographenes (20-35 kcal·mol−1) Barriers beyond 20 kcal·mol−1 usually allow the isolation and manipulation of enantiopure compounds. However, the racemization process occurs over time and the chiroptical properties are not stable enough for most of the applications. Although the introduction of non-hexagonal rings usually leads to a high flexibility of the structure and therefore a low isomerization barrier, one of the possibilities to increase the rigidity is to fuse rings smaller and larger than hexagons. This is the case of the monkey-saddle shape nanographene 2.7 containing three five-membered rings and three eight-membered rings fused to a central benzene (Figure 2.6). The combination of these topological defects leads to an isomerization barrier of 24.8 kcal·mol−1.113 Alternatively, negatively curved nanographenes in which the non-hexagonal ring is embedded in a helicene-like topological defect exhibit higher stereochemical rigidity. As shown in Figure 2.6, nanographene 2.8 has two seven-membered rings, one forming a hexa[7]circulene and the other one embedded forming a [5]helicene, which results on a racemization barrier of 25.4 kcal·mol−1.114 Figure 2.6. Isolable chiral nanographenes containing non-hexagonal rings and helical topological defects. Structures containing [5]helicenes are expected to be in this subclass considering the isomerization barrier of [5]helicene is 24.6 kcal·mol−1, however, strategies as the functionalization of the terminal rings (fjord region) increase the racemization barrier. 111 X. Gu, H. Li, B. Shan, Z. Liu, Q. Miao, Org. Lett. 2017, 19, 2246. 112 S. H. Pun, Y. Wang, M. Chu, C. K. Chan, Y. Li, Z. Liu, Q. Miao, J. Am. Chem. Soc. 2019, 141, 9680. 113 T. Kirschbaum, F. Rominger and M. Mastalerz, Angew. Chem., Int. Ed. 2020, 59, 270. 114 Z. Qiu, S. Asako, Y. Hu, C.-W. Ju, T. Liu, L. Rondin, D. Schollmeyer, J.-S. Lauret, K. Müllen, A. Narita, J. Am. Chem. Soc. 2020, 142, 14814. Chapter 2 67 For instance, nanographene 2.9 has a racemization barrier of 33 kcal·mol−1 arising from the [5]helicene functionalize with two bulky tert-butyl groups (Figure 2.6).115 2.1.2.4 Stereochemically Rigid Nanographenes (>35 kcal·mol−1) At 423 K, an enantiomerically pure sample racemizes with a half-time life of τ1/2= 24 h when the isomerization barrier is 35 kcal·mol−1. The chiral properties are stable enough for chiroptical applications. The methodologies above described and the presence of helicenes larger than [6]helicenes ensure the stereochemical rigidity. As describe above, rings larger than six members embedded in a helicene-like structure increases the rigidity. Nanographene 2.10 contains an eight-membered ring involved in a [5]helicene-like structure (Figure 2.7). The racemization energy increase is such that the measurable lower limit of the barrier resulted in 38.3 kcal·mol−1.116 Moreover, the functionalization of the fjord region of [5]helicenes with bulky groups was sufficient to obtain the stereochemically rigid nanographene 2.11. Additionally, the helicity of the substituted [5]helicenes induce the twist of the structure increasing the barrier to 38.4 kcal·mol−1 (Figure 2.7).117 Figure 2.7. Stereochemically rigid nanographenes containing non-hexagonal rings and helical topological defects. The stereochemical rigidity of helicene containing nanographenes depends on the number of ortho-fused rings. Taking into account that the isomerization barrier of [6]helicene is 36.2 kcal·mol−1, it can be expected that the introduction of longer helicenes will lead to stereochemically rigid structures. For instance, nanographene 2.12 consists of two [6]helicene and two [7]helicene substructures, a combination that leads to an isomerization barrier of 58.6 kcal·mol−1 (Figure 2.7).118 115 C. M. Cruz, I. R. Márquez, I. F. A. Mariz, V. Blanco, C. Sánchez-Sánchez, J. M. Sobrado, J. A. Martín- Gago, J. M. Cuerva, E. Maçôas, A. G. Campaña, Chem. Sci. 2018, 9, 3917. 116 M. A. Medel, R. Tapia, V. Blanco, D. Miguel, S. P. Morcillo, A. G. Campaña, Angew. Chem., Int. Ed. 2021, 60, 6094. 117 X. Xu, R. Munoz-Marmol, S. Vasylevskyi, A. Villa, G. Folpini, F. Scotognella, G. M. Paterno, A. Narita, Angew. Chem. Int. Ed. 2023, 62, e202218350. 118 S. H. Pun, K. M. Cheung, D. Yang, H. Chen, Y. Wang, S. V. Kershaw, Q. Miao, Angew. Chem. Int. Ed. 2022, 61, e202113203. Chapter 2 68 2.1.3 Alternative Approaches for Chirality in PAHs As shown in the previous section, the design of chiral nanographenes is mainly based on the introduction of topological defects that give rise to Gaussian curvature or helicity. Only a few examples of alternative approaches to chiral polyaromatic hydrocarbons can be found in the literature. For instance, Itami and coworkers reported a cyclic chiral polyarene, infinitene 2.13, that consists of two fused [8]helicene units acquiring the shape of the infinity symbol.119 As shown in Figure 2.8, the structure, which could be considered a helically twisted analogue of a saddle- shaped [12]circulene, fulfills the symmetry conditions as well as the stereochemical rigidity, so it is chiral. A larger 8-shaped helical cyclic stereochemically rigid nanobelt 2.14 was described by Wu et al. The isomerization barrier was estimated to be higher than 82.5 kcal·mol−1 (Figure 2.8).120 More recently Wu and coworkers reported a large twisted Möbius carbon nanobelt 2.15 (Figure 2.8). The Möbius molecules are also inherently chiral, in this case the band-like structure is twisted three times, which involves three chiral axis leading to two stable enantiomers (P,P,P or M,M,M).121 In addition, untwisted chiral nanobelts as 2.16122 can be considered "molecular nanotube fragments", whose chirality steams from the bending vectors of the hypothetically open planar structure. Figure 2.8. Chiral cyclic polyaromatic hydrocarbons, nanobelts. An almost unexplored approach to chiral nanographenes is the design of atropisomers. As shown in Figure 2.9, the hindered rotation around the single bond (in purple) connecting the aromatic substituents leads to enantiomers when the ortho-substituents are different two by two. Thus, the cyclic structure 2.17 has three chiral axes along the simple bonds (highlighted in purple), leading to three possible pairs of enantiomers of a stereochemically rigid structure. However, the reported structures are the 119 M. Krzeszewski, H. Ito, K. Itami, J. Am. Chem. Soc. 2022, 144, 862. 120 W. Fan, T. Matsuno, Y. Han, X. Wang, Q. Zhou, H. Isobe, J. Wu, J. Am. Chem. Soc. 2021, 143, 15924. 121 W. Fan, T. M. Fukunaga, S. Wu, Y. Han, Q. Zhou, J. Wang, Z. Li, X. Hou, H. Wei, Y. Ni, H. Isobe, J. Wu, Nat. Synth. 2023, 2, 880. 122 K. Yin Cheung, S. Gui, C. Deng, H. Liang, Z. Xia, Z. Liu, L. Chi, Q. Miao, Chem 2019, 5, 838. Chapter 2 69 enantiomers (R,R,R or S,S,S) resulting from enantiopure starting materials (R or S BINOL).123 Figure 2.9. Biphenyl chirality of a cyclic polyaromatic hydrocarbon and a molecular nanographene. To the best of our knowledge, the only nanographene with atropisomerism described in the literature was reported in 2023 by An et al. The structure 2.18 consists of two graphitized units containing a heptagonal ring forming a cyclic ether. The graphitized fragments are connected by a single bond through which the rotation is restricted. Thus, the symmetry and stereochemical rigidity requirements are fulfilled, considering that the experimentally determined racemization barrier was found to be 35 kcal·mol−1 at 443 K (Figure 2.9).124 123 Y. Nojima, M. Hasegawa, N. Hara, Y. Imai, Y. Mazaki, Chem. Commun. 2019, 55, 2749. 124 S. Li, R. Li, Y.-K. Zhang, S. Wang, B. Ma, B. Zhang, P. An, Chem. Sci. 2023, 14, 3286. 70 Chapter 2 71 2.2 Objectives This chapter is focused on the search for new strategies for the synthesis of molecular nanographenes with atropisomerism. For this purpose, a family of orthogonally connected nanographenes based on two graphitized units (dibenzo[fg,ij]phenanthro- [9,10,1,2,3-pqrst]pentaphene, DBPP) attached to a central core was designed. Although the first attempt was to study anthracene as the central core, structures 1 (not obtained) and 13 described in the previous Chapter 1, the chirality of these systems has been discussed by using simplified structures, replacing the anthracene (or octafluoroanthracene) core by a 1,4-substituted tetrafluorobenzene considering its easier functionalization and accessibility. The work plan consisted on the synthesis of the symmetric derivative and subsequent modification of the substituents, in the graphitized moieties and the central core, to obtain the atropisomers. However, upon synthesizing the symmetric derivative it was observed how the steric hindrance entailed the organization of the graphitized units, giving rise to the unexpected chirality of the highly symmetrically substituted nanographene (1,4-substituted tetrafluorobenzene connecting the DBPP units substituted with five tert-butyl groups). Furthermore, racemization was observed while attempting to isolate the enantiomers. Thus, the stereochemical rigidity was lower than expected. To determine the isomerization barrier, two asymmetric derivatives were prepared by exchanging two tert-butyl groups for -OMe or -H, and the isomerization process was evaluated by NMR experiments. Finally, the study of the optoelectronic properties was carried out by cyclic voltammetry, absorption, and emission measurements. 72 Chapter 2 73 2.3 Results and Discussion 2.3.1 Synthesis of Helically Arranged Nanographenes In this section, the synthesis and characterization of a new family of helically arranged nanographenes based on two graphitized units (dibenzo[fg,ij]phenanthro-[9,10,1,2,3- pqrst]pentaphene, DBPP) connected through a 1,4-substituted tetrafluorobenzene ring are described. As detailed in Scheme 2.1, the synthetic methodology starts with a double palladium-catalyzed Sonogashira coupling between two equivalents of a suitably para-substituted (R= tert-butyl, methoxy or hydrogen) arylacetylene 3a-c and 1,4-dibromo-2,3,5,6-tetrafluorobenzene (14). The resulting bis[aryl(ethynyl)] tetrafluorobenzenes 15a-c undergo a double Diels-Alder cycloaddition with cyclopentadienone 5 to afford the corresponding polyarene structures 16a-c. The final step is the graphitization by the Scholl reaction of the respective polyphenylene in presence of DDQ and triflic acid, thereby effectively yielding helically arranged nanographenes 17a-c. Scheme 2.1. Synthesis of tetrafluorobenzene-based helically arranged nanographenes. Chapter 2 74 As described in the previous Chapter 1, the formation of trityl cation intermediates during the Scholl reaction leads to the formation of spirocycles between the graphitized units and the central core (Figure 1.13). However, the substitution of the central anthracene with eight electron withdrawing fluorine atoms conducts the reaction to the formation of the expected helically arranged nanographene 13. The fluorine atoms in the above described nanographenes 17a-c have a double function, destabilizing the hypothetical carbocation and preventing the six-membered ring cyclization between the graphitized DBPP units and the central ring. Furthermore, the introduction of fluorine atoms into polyarene structures has proven to have no negative impact on the Scholl oxidation leading to unexpected rearrangements.125 The characterization by 1H-NMR, 13C-NMR, 19F-NMR, FT-IR, and high-resolution mass spectrometry confirmed the structures. The NMR spectra corresponding to nanographene 17a reveals its high symmetry. As shown in Figure 2.10, the spin systems, integrals, and multiplicity are the same as those for octafluoroanthracene- based nanographene 13 by 1H-NMR: one doublet of doublets at 7.63 ppm (3JH-H= 8.77 Hz and 4JH-H= 1.90 Hz), a doublet at 8.26 ppm (3JH-H= 8.77 Hz), one doublet at 8.81 ppm (4JH-H= 1.90 Hz), two doublets at 9.01 and 9.11 ppm (4JH-H= 1.65 Hz), one singlet at 9.16 ppm, and tree singlets at 1.82, 1.76 and 1.39 ppm corresponding to the tert- butyl groups with relative intensities 1:2:2. The 19F-NMR shows one singlet corresponding to the four magnetically equivalent fluorine atoms in the central benzene ring. Figure 2.10. 1H-NMR spectrum (left) showing the relevant regions, assigned structure in the middle, and 19F-NMR spectrum (right). 125 S. Mörsel, R. Kellner, A. Hirsch, Eur. J. Org. Chem. 2023, 26, e202300299. 1H, 17a 19F, 17a Chapter 2 75 2.3.2 Unexpected Chirality The structure was unequivocally resolved by single-crystal X-ray diffraction, shown in Figure 2.11. Nanographene 17a presents a helical arrangement of the moieties. The steric hindrance between the DBPPs themselves, and with the central tetrafluorobenzene ring, and the possible rotation around the single bonds connecting the fragments cause the structure to acquire a helical organization around a C2 axis (shown in yellow, Figure 2.11a), as previously observed in the octafluoroanthracene- based nanographene 13 (Chapter 1). To identify the helicity, numbers have been assigned in increasing order to the structural units considering the view along the C2 axis (Figure 2.11b). Starting from the closest to the observer (1) and passing through the central benzene unit (2) to the furthest (3). Figure 2.11. a) Front view of nanographene 17a. b) Lateral view of nanographene 17a showing the helical arrangement. C2 axis represented in yellow. Detailed observation of the nanographene in the crystal structure showed that the DBPP fragments exhibit an unexpected curvature, Figure 2.12a. Density functional theory (DFT) calculations on the model system 17aM, in which the tert-butyl groups were replaced by methyl groups, revealed the occurrence of stabilizing non-covalent C−H···π interactions (green surfaces, Figure 2.12b). The interactions between the closest peripheral rings of the two DBPPs and between the tetrafluorobenzene ring with the closest C–H bonds in the graphitized fragments, lock the peripheral rings orthogonally (as represented with the planes in orange and purple in Figure 2.12a) inducing the curvature of the DBPP moieties. According to the second order perturbation theory (SOPT) of the natural bond order (NBO) method, the attractive interactions result from the π(C=C)→σ*(C−H) and σ(C−H)→π*(C=C) electronic delocalization, whose associated stabilization energies, ΔE(2), amount to −1.04 and −0.68 kcal·mol-1, respectively. (a) (b) C2 C2 1 2 3 Chapter 2 76 Figure 2.12. a) Orthogonal disposition of the interacting peripheral rings in the DBPP fragments. b) Stabilizing C–H···π interactions (surfaces in green) that set the structure geometry. Therefore, the steric hindrance between the DBPPs prevents their coplanarity. Interestingly, this deviation from planarity entails the bending of the DBPPs, and the structure is fixed, so that the three fragments are arranged in a helical fashion. Figure 2.13a shows the planes containing the central rings of each fragment (the flatter area). The plane represented in red contains the central part of the DBPP fragment closest to the observer, the plane in green contains the tetrafluorobenzene ring (located in the center), and finally the orange plane contains the central rings of the fragment farthest from the observer. The helical arrangement of the moieties means that the structure contains neither symmetry planes nor inversion centers. The symmetry operations in nanographene 17a are three orthogonal C2 rotational axis, therefore, the structure is framed in the chiral space-group of symmetry D2. Evaluating the packing in the crystal structure it was noticed that two non- superposable specular images of the nanographene coexist in a 1:1 relation. The symmetrical substitution of the central ring does not allow the application of the Cahn- Ingold-Prelog priority rules. As represented in Figure 2.13b, the numbers assignation to the fragments, from the closest to the observer to the farthest, results in a clockwise (P isomer) or counterclockwise (M isomer) helicity. Figure 2.13. a) Lateral view of nanographene 17a representing the planes containing each unit. In red, DBPP unit closest to the observer, in green, tetrafluorobenzene, and in orange, farthest DBPP unit. b) P (clockwise helicity) and M (counterclockwise helicity) enantiomers of nanographene 17a coexisting in the crystal structure. 90º (a) 17aM17a (b) (b) 1 2 3 1 2 3M P (a) Chapter 2 77 Enantiomers of nanographene 17a were resolved by chiral HPLC (Appendix 2, HPLC resolution). However, they could not be isolated as enantiopure samples for the determination of chiral properties due to racemization at room temperature. For the identification of the isomerization barrier responsible for the racemization, it was required to understand the molecular dynamics. Therefore, two asymmetric derivatives were prepared by changing one tert-butyl group on each DBPP by a methoxy group or hydrogen atom, as shown in Scheme 2.1, providing the nanographene derivatives 17b and 17c, respectively. 2.3.3 Molecular Dynamics: Asymmetric Derivatives 17b,c Nanographenes 17b and 17c were obtained as 1:1 anti/syn diastereomers mixture when the Scholl reaction was performed at 0 ºC. DFT calculations supported this result indicating that the diastereomers are nearly degenerate. 1H-NMR spectra of nanographenes 17b and 17c show the duplication of all signals, e.g. two singlets are observed at (3.95 and 3.93 ppm, Figure 2.14a) corresponding to the methoxy groups on each isomer of 17b. Both 19F-NMR spectra showed the presence of two different fluorine atoms corresponding to the anti and syn isomers of each compound (Figure 2.14b). Figure 2.14. a) 1H-NMR spectra of nanographene 17b. b) 19F-NMR spectra of nanographenes 17b and 17c. The stereoisomers are anti, if the substituents (-OMe or H) are facing different sides of the tetrafluorobenzene, or syn if they are facing the same plane of the central core, Figure 2.15a. Thus, these derivatives are obtained as a four-isomers mixture, a pair of diastereomers with their corresponding pair of enantiomers. In Figure 2.15b, the lateral views of each isomer are shown and, to facilitate the isomers assignation, simplified models are represented. The dark grey sticks represent the DBPP closer to the observer, the tetrafluorobenzene ring is represented as a green line with green dots in the ends representing the fluorine atoms, and the light grey sticks represent the DBPP farther from the observer. Finally, the OMe (17b) or H (17c) substituents are shown as purple dots. (a) 1H, anti/syn 17b Anti/syn 17b Anti/syn 17c (b) 19F Chapter 2 78 Figure 2.15. a) Front view of anti/syn 17b (purple dot: OMe) or 17c (purple dot: H). b) Lateral view of the anti/syn diastereomers, assignation of the M/P enantiomers. Simplified models are used for better visualization. c) Isomerization by rotation of the central ring: anti→syn. d) Isomerization by crossing of the DBPP units: M→P. (a) Front view Anti Syn (b) Lateral view Anti Syn Diastereomers (c) 90º Rotation of the central C6F4 f(t): Anti / Syn isomerization 1 2 3 1 2 3 1 2 3 1 2 3 P M M P Anti 1 2 3 t0 tSyn 1 2 3 P (d) Rotation of the DBPP moieties f(t): Racemization M Anti 1 2 3 t0 Anti M 1 2 3P t E n a n ti o m e rs 1 2 3 P 1 2 3 M 1 2 3 M 1 2 P 3 – OMe (17b) –H (17c) 90º 90º 90º Chapter 2 79 As previously explained, the P enantiomer is that in which the fragments are organized clockwise around the chiral axis, and M in which the fragments are organized counterclockwise. To identify the isomerization barrier (previously revealed by the racemization of nanographene 17a), two possible scenarios should be considered. On the one hand, a 90º rotation of the tetrafluorobenzene core, which leads to isomerization between diastereomers (e.g. anti M to syn P, Figure 2.15c). On the other hand, the rotation of the DBPP moieties with respect to each other (crossing), entailing racemization, interconversion between enantiomers (e.g. anti M to anti P, Figure 2.15d). The isomerization barrier was determined by 1H-NMR experiments. To avoid possible electronic effects resulting from the methoxy groups in derivative 17b, derivative 17c, in which two tert-butyl groups have been removed, was used. As previously mentioned, compound 17c was isolated as 1:1 anti/syn mixture when the Scholl reaction was performed with DDQ and triflic acid at 0 ºC. However, at −65 ºC, the Scholl oxidation of polyphenylene 16c yielded a 70:30 anti/syn isomeric mixture (anti and syn isomers assignations by NMR in Appendix 5) with noncomplete conversion. After workup to remove the DDQ and triflic acid, the mixture was warmed at 40 ºC and monitored by 1H-NMR recording a spectrum every 10 minutes. After assigning the signals to each isomer by NMR experiments, it was decided to monitor four doublets that were sufficiently separated. The doublets at 8.31 and 8.14 ppm belonging to the anti isomer, and the doublets at 8.27 and 8.22 to the syn isomer. As shown in Figure 2.16, at the beginning of the experiment (t0) the anti/syn ratio is 70:30 and as time progresses the integrations evolved until both isomers are at a 50:50 ratio after 15 000 seconds. Figure 2.16. Isomerization process monitored by performing a 1H-NMR spectrum every 10 minutes at 40 ºC. Anti Syn t0 15 000 s Chapter 2 80 The kinetic constants were determined according to the integrals’ variations. Since the concentration of both anti/syn diastereomers is the same at the equilibrium, the constants kf and kb are the same (Figure 2.17a). The representation of the ln[(XH,8.31−Xeq)/(X0−Xeq)] (XH,8.31, molar fraction of the proton at 8.31 ppm depending on time; Xeq, molar fraction of the proton at 8.31 ppm once reached the equilibrium, X0, molar fraction of the proton at 8.31ppm at the beginning of the experiment) vs time provided a slope of −2.15·10-4 s-1 with a good fitting (r2= 0.9849), Figure 2.17b. According to the kinetic equation, the slope corresponds to –(kf+kb), then kf = kb = 1.08·10−4 s−1. If these kinetic constants value is introduced to Eyring’s equation, k=((κ·kB·T)/h)·e^(−ΔG≠/(RT)), the isomerization barrier at 40 ºC is Δ𝐺⧧= 24.6 kcal·mol-1, which corresponds to a half-time life of τ1/2 = 107 min. Figure 2.17. a) Isomerization process monitored by 1H-NMR. b) Variation of the molar fraction of anti/syn isomers with time (left) and fitting of ln[(XH,8.31−Xeq)/(X0−Xeq)] vs t (right). Since NMR was used for the determination of the isomerization barrier from anti 17c to syn 17c, only differentiation between diastereomers is possible, and not between enantiomers (nuclei present an equivalent chemical and magnetic environment). Then, it can be concluded that the isomerization process involves a 90º rotation of the tetrafluorobenzene ring. Which results in interconversion between diastereomers as previously illustrated in Figure 2.15c. kf kb At eq. [anti]=[syn]→kf =kb Anti Syn (a) (b) 0 2500 5000 7500 10000 12500 -3.0 -2.8 -2.6 -2.4 -2.2 -2.0 -1.8 -1.6 -1.4 -1.2 -1.0 -0.8 -0.6 -0.4 -0.2 0.0 Linear fitting Slope -2.15·10 -4 s -1 r 2 0.9849 ln [( X -X e q )/ (X 0 -X e q )] Time (s) Equation y = a + b*x Weight No Weighting Residual Sum of Squares 0.19798 Pearson's r -0.9928 Adj. R-Square 0.9849 Value F Intercept 0.12129 Slope -2.14755E-4 0 3000 6000 9000 12000 15000 0.30 0.35 0.40 0.45 0.50 0.55 0.60 0.65 0.70 Xanti X syn M o la r fr a c ti o n , X Time (s) Chapter 2 81 2.3.4 No C−H ··· π Interactions, No Helicity, No Chirality Considering the backbone of the central rings (highlighted in blue, Figure 2.18), this nanographene, a priori, may be considered a π-extended para-terphenyl. Depending on the substitution pattern of the central ring, para-terphenyls can be classified in four types.126 The asymmetric substitution on the central benzene entails the absence of mirror planes and inversion elements, thus different number of isomers are formed depending on the remaining symmetry elements. However, when all substituents are the same in the central ring (Type 4, Figure 2.18), only one isomer is expected, reflection symmetry elements, mirror plane containing the lateral benzene rings and mirror plane perpendicular to the central ring, annuls the presence of chirality. However, the symmetry planes present on Type 4 para-terphenyls are not possible symmetry operations for nanographene 17a. The structure presents two enantiomers due to the C–H···π interactions that set the structure with a helical arrangement of the fragments DBPP-tetrafluorobenzene-DBPP. Thus, inducing atropisomerism-like chirality in a highly symmetric molecule. Figure 2.18. Considering nanographene 17a a π-extended terphenyl and relation with the terphenyl types. Moreover, the non-graphitized structures of polyarenes 16b and 16c present the symmetry of type 4 terphenyls. The lack of the previously described C−H···π interactions (Figure 2.12) due to free bonds rotations means the presence of mirror planes and, therefore, no chirality. As shown in Figure 2.19, the NMR spectra of the polyarene asymmetric derivatives 16b and 16c present signals corresponding to only one isomer (e.g. one singlet corresponding to the methoxy group, four singlets corresponding to the tert-butyl groups and one singlet in the 19F-NMR spectra). 126 R. Adams, H. C. Yuan, Chem. Rev. 1933, 12, 261 X Chapter 2 82 Figure 2.19. a) 1H-NMR spectrum of polyarene derivative 16b. b) 19F-NMR spectra of polyphenylenes 16b and 16c. Only one compound is observed, no isomers. In addition, performing the Scholl oxidation of polyarene 16c at −78 ºC provided the partially graphitized derivative 17c’ (Figure 2.20), for which only one isomer is observed by NMR experiments, e.g. four singlets corresponding to the tert-butyl groups and one singlet in 19F-NMR, as shown in Figure 2.20 (assignation by NMR in Appendix 5), therefore it can be considered a π-extended type 4 terphenyl (Figure 2.18). The free rotation of the fragments around the simple bonds entails the lack of stereogenic elements. The most stable conformation calculated by DFT 17c’M shows a non-chiral orthogonal arrangement, which contrasts to the helical arrangement of chiral nanographene 17c. Figure 2.20. a) 1H and 19F-NMR spectra of partially graphitized non-chiral derivative 17c’. b) DFT modelized structure 17c’M showing non-chiral orthogonal arrangement of the moieties. 16b 16c (a) 1H, 16b (b) 19F 17c’M (a) 1H, 17c’ 19F, 17c’ (b) Chapter 2 83 2.3.5 Electrochemical and Photophysical Properties The electrochemical properties of nanographenes 17a-c were explored by cyclic voltammetry in a toluene/acetonitrile 4:1 mixture using tetrabutylammonium hexafluorophosphate as supporting electrolyte, glassy carbon as working electrode, a platinum wire as counter electrode, and Ag/AgNO3 as reference electrode, at room temperature. Molecular nanographenes 17a-c present non-reversible oxidation and reduction waves with rather extreme potentials for the technique, finding the main limitation in the solvent window (oxidation and reduction of the solvent). After testing several measurement conditions, the cyclic voltammograms spectra represented in Figure 2.21 show the oxidation and reduction potentials, except for the oxidation of nanographenes 17a and 17c, for which the square wave voltammetry curves are illustrated with dashed lines. Given the low reversibility of most of the observed waves, the values of the maxima observed by square wave voltammetry are reported. Figure 2.21. Cyclic voltammograms corresponding to tBu-HBC, and nanographenes 17a-c vs Fc/Fc+ in a 1 M solution of tetrabutylammonium hexafluorophosphate in toluene/acetonitrile 4:1. With dashed lines are represented the square wave measurements. The oxidation and reduction potentials vs. Fc/Fc+ of the new nanographenes and tBu- HBC are summarized in Table 2.1. Compared to tBu-HBC, these nanographenes are poorer electron donor and acceptors, with first reduction waves at E1 red= –2.42 V (17a), E1 red= –2.39 V (17b), and E1 red= –2.34 V (17c) vs. E1 red= –2.26 V (tBu-HBC), and oxidation waves at E1 ox= 0.92 V (17a), E1 ox= 0.94 V (17b), and E1 ox= 0.97 V (17c) vs. E1 ox= 0.80 V (tBu-HBC). This behavior suggests that the conjugation between the three fragments (DBPPs and tetrafluorobenzene core) is hampered because of the non- coplanarity. Therefore, the π-conjugation and extension is lower in the helical nanographenes than in tBu-HBC. These observations nicely correlate with the DFT- Chapter 2 84 computed energy of the corresponding HOMO (orbital from which the electron is released): −5.54 eV (17aM) < −5.21 eV (tBu-HBC), thus showing that a more stabilized HOMO (i.e., more negative) is translated into a higher oxidation potential. Table 2.1. Redox potentials of tBu-HBC, 17a, 17b and 17c vs Fc/Fc+. Compound E1 red (V) E1 ox (V) tBu-HBC –2.26 0.80 17a –2.42 0.92 anti/syn 17b –2.39 0.94 anti/syn 17c –2.34 0.97 The optical properties of nanographenes 17a-c were evaluated by UV-vis absorption and emission measurements (Figure 2.22). The absorption spectra of this family of molecular nanographenes show similar shapes and energies (Figure 2.22, solid lines), with three sharp intense absorption bands in the UV region (17a: 317, 356, and 378 nm; 17b: 320, 357, and 380 nm; 17c: 318, 354, and 377 nm) and three weak bands in the visible region (17a: 395, 419, and 444 nm; 17b: 397, 419, and 445 nm; 17c: 395, 419, and 444 nm). Comparing with tBu-HBC, the shape is very similar, presenting three sharp bands in the UV region (344, 360, and 390 nm) and three weak bands in the visible region (439, 441, and 443 nm). The blue shifting of the bands of nanographenes 17 vs. tBu-HBC agrees with the electrochemical properties previously described, since the structures are less conjugated, which also involves the increment of the optical band gap (calculated in the intersection between the absorption and emission spectra, 17a E0-0= 2.79 eV, 17b E0-0= 2.78 eV, 17c E0-0= 2.79 eV, tBu-HBC E0-0= 2.65 eV). Figure 2.22. Absorption (solid line) and emission (dashed line) spectra of nanographenes 17. Chapter 2 85 To determine the nature of the transitions associated to the obtained UV-vis spectra, time-dependent (TD)DFT calculations were carried out on the model 17aM. The theoretical data nicely reproduce the occurrence of the bands, the more intense band at 395 nm (λcalc= 410 nm, f= 0.18) is assigned to a transition from HOMO−2 to LUMO. The two bands at 419 and 444 nm (λcalc= 418 and 420 nm, respectively), show a rather low oscillator strength (f= 0.028 and 0.020, respectively), which agrees with rather low ε experimentally observed. These last transitions result from a one-electron transition from the nearly degenerate HOMO and HOMO−1 (π-molecular orbitals delocalized in both DBPP moieties with no measurable coefficient in the central tetrafluorobenzene ring, Figure 2.23) to the LUMO, respectively. It is worth to mention, that the LUMO is delocalized along three fragments due to the presence of twisted π-orbitals connecting the central ring with the DBPPs. The HOMO−2 also exhibits coefficients in the central ring, which confirms π-communication in the nanographene despite the lack of coplanarity between the fragments. Figure 2.23. Computed molecular orbitals for 17aM involved in the main UV-vis absorptions. The emission spectra of these nanographenes are also very similar, showing three bands in the visible region (17a: 446, 474, and 505 nm; 17b: 448, 476, and 507 nm; 17c: 446, 473, and 504 nm). In these spectra de bands are also blue shifted if compared to tBu-HBC (absorption and emission maxima are collected in Table 2.2). In addition, the optical energy band gaps were calculated by the intersection of the absorption and emission spectra resulting of 2.79 eV (17a), 2.78 eV (17b), and 2.79 eV (17c). These energy gaps are larger than that of tBu-HBC (2.69 eV), because of the lower π-extension of helically arranged nanographenes. Chapter 2 86 Table 2.2. Absorption and emission bands of tBu-HBC, 17a, 17b and 17c. Compound Absorption λabs max (nm) Emission λem max (nm) tBu-HBC 344, 360, 390, 439, 441, 445 493, 519, 553 17a 317, 356, 378, 395, 419, 444 446, 474, 505 Anti/syn 17b 320, 357, 380, 397, 419, 445 448, 476, 507 Anti/syn 17c 318, 354, 377, 395, 419, 444 446, 519, 553 2.3.6 Photophysical Properties of Helically Arranged Nanographenes As previously mentioned in Chapter 1.3.5, the emission spectra of octafluoroanthracene-based nanographene 13 shows a dual emission that could stem from the apparently individual emission of the chromophores, octafluoroanthracene and DBPP moieties. This observation is based on the similarities shown by the emission spectra of nanographene 17a and precursor 12 with the emission spectrum of nanographene 13, whose shape suggests the combination of both emission spectra. As shown in Figure 2.24, the emission bands for nanographene 13 at lower wavelengths could be associated to the emission of the DBPP moieties (similar in shape to the emission spectrum of 17a), whereas the more intense broad emission bands at longer wavelengths may be related to the octafuoroanthracene core (similar in shape to the emission spectrum of 12). Figure 2.24. Emission spectra comparison between compound 12 and helically arranged nanographenes 13 and 17a. Chapter 2 87 2.4 Conclusions The synthesis and characterization of a new series of helically arranged nanographenes consisting of two DBPP moieties covalently connected to a central 1,4-substituted tetrafluorobenzene core has been described. Single-crystal X-ray diffraction of nanographene 17a showed a deviation from planarity of the DBPP moieties. DFT calculations reproduce the crystalline structure and revealed the formation of stabilizing C−H···π interactions that set the structure and support the unexpected bending of the DBPPs. The resulting helical arrangement of the moieties around the chiral axis leads to the formation of two enantiomers for the highly symmetrical nanographene 17a. The enantiomers were resolved by chiral HPLC; however, the isolation was not successful due to racemization at room temperature. The isomerization process was evaluated by NMR experiments of the asymmetrically substituted derivative 17c. The replacement of two tert-butyl groups leads to a new pair of nearly degenerated diastereomers (anti/syn) with a rotational barrier of Δ𝐺⧧= 24.6 kcal·mol−1 at 40 ºC, and a half-time life of τ1/2= 107 min. The absence of the C−H···π interactions entails the absence of chirality, as proven by the achiral polyarene precursors (16b and 16c) and the also achiral partially graphitized nanographene 17c’. Structures in which the helical organization does not occur and, therefore, can be classified as type 4 π-extended para-terphenyls. The electrochemical properties were evaluated by cyclic voltammetry and square wave voltammetry. Helically arranged nanographenes 17a-c are poorer electron donor and acceptors than tBu-HBC, with first wave reduction values of E1 red < −2.42 V (17a), vs. E1 red= −2.26 V (tBu-HBC), and first wave oxidation values of E1 ox > 0.92 V (17a), vs. E1 ox= 0.80 V (tBu-HBC). The absorption and emission spectroscopic properties confirmed the same trend, the maxima are blue shifted in comparison with those of tBu-HBC, which is related to the lower π-extension of nanographenes 17a-c vs. tBu- HBC. However, DFT calculations interestingly showed that the LUMO orbital is delocalized along the entire molecule due to the presence of twisted π-orbitals connecting the central ring with the DBPP fragments, thus allowing the electronic communication along the entire molecule. Chapter 3 Chapter 3 91 3 Chapter 3. 3.1 Introduction to Bilayer Nanographenes As described in Chapter 1.1, the wide variety of organic chemistry reactions allows the stepwise synthesis of molecular nanographenes with atomic control and, therefore, perfectly defined properties. The variety of structures reported in the literature gives evidence of the numerous possibilities for modifying the size and shape of these graphene molecular fragments. The main approaches are the variation of the π- extension, and the topological modification of the hexagonal networks, allowing the induction of inherent chirality, as detailed in Chapter 2.1. However, having the synthetic tools, imagination is the limit and, more recently, special interest has been focused on the design of bilayer nanographenes. Resembling the multilayer structure of graphite, planar nanographenes tend to aggregate by π-π intermolecular interactions. Taking advantage of this behavior, in 1999 Müllen et al. reported the first supramolecular aggregates of HBC derivatives forming ordered columnar structures.127 Thus, in the early 2000’s a great development of ordered HBC-based liquid crystals128 took place due to its interest in organic electronics.129 However, it has not been until the last few years that discrete persistent bilayer and multilayer systems could be prepared in a controlled manner, i. e. with a fixed number of overlapping layers stacked by π-π interactions. Depending on the binding forces involved in the bilayer formation, three different approaches have been explored, (i) purely supramolecular interactions providing van der Waals molecular nanographenes, (ii) bilayers from fused radicals, and (iii) covalently connected bilayers. 3.1.1 Van der Waals Bilayer Molecular Nanographenes The design of discrete, persistent bilayers by this approach can be challenging due to the labile nature of van der Waals interactions. In addition, the monomer must be design in a way that the supramolecular polymerization is hindered. To the best of our knowledge, the first van der Waals persistent molecular bilayers were reported in 2019 by Tan and coworkers.130 As shown in Figure 3.1, the structure 3.1 is a dimer of a planar nanographene functionalized with 2,6-dimethylphenyl substituents in the periphery. The main purpose of the bulky peripheral groups is to provide solubility to nanographenes preventing the formation of multiple aggregates by π-π stacking. 127 A. Fechtenkötter, K. Saalwächter, M. A. Harbison, K. Müllen, H. Wolfgang Spiess, Angew. Chem. Int. Ed. 1999, 38, 3039. 128 T. Wöhrle, I. Wurzbach, J. Kirres, A. Kostidou, N. Kapernaum, J. Litterscheidt, J. Christian Haenle, P. Staffeld, A. Baro, F. Giesselmann, S. Laschat, Chem. Rev. 2016, 116, 1139. 129 J. P. Hill, W. Jin, A. Kosaka, T. Fukushima, H. Ichihara, T. Shimomura, K. Ito, T. Hashizume, N. Ishii, T. Aida, Science 2004, 304, 1481. 130 X.-J. Zhao, H. Hou, X.-T. Fan, Y. Wang, Y.-M. Liu, C. Tang, S.-H. Liu, P.-P. Ding, J. Cheng, D.-H. Lin, C. Wang, Y. Yang, Y.-Z. Tan, Nat. Commun. 2019, 10, 3057. Chapter 3 92 However, the suitable design of the monomer allows large π-π interactions between two layers. The interlocking between the orthogonal peripheral groups prevents the formation of larger aggregates providing a persistent bilayer structure that only dissociates into the corresponding monomers by laser ablation. The strong van der Waals interactions entail the formation of excited states with different energies to those of the degenerate states of the monomers −Davydov splitting−. Transient absorption and time-resolved photoluminescence spectroscopic measurements showed significant differences between the lifetimes of the bright and dark Davydov states resulting from the dimerization. Figure 3.1. Persistent van der Waals bilayer and trilayer nanographenes. The successfully proven strategy to prevent the polymerization while enhancing the π-π interactions by the design of suitably substituted π-extended structures was followed by Würthner et al. 131 In 2022 they described the controlled formation of trilayer 3.2 by alternating π-π stacking between different monomers, a trisimide nanographene and two functionalized HBC monomers. As illustrated in Figure 3.1, the trisimide nanographene can encompass exactly two equivalents of a same symmetry hexabenzocoronene (in yellow) to form the persistent trilayer 3.2. Interestingly, the formation of the van der Waals structure protects the trisimide nanographene π-surface from oxygen quenching, improving the fluorescence under ambient conditions (ΦFL= 12.5% for the trisimide nanographene vs. ΦFL= 21.2% for the trilayer). More recently, Mastalerz and coworkers have described a new family of twisted chiral acenes that undergo an enthalpy driven and solvent dependent dimerization process leading to van der Waals bilayers.132 The ribbon monomers forming the twisted bilayer 3.3 (Figure 3.2), consist of 25 linearly annulated benzene rings. The peripheral 131 B. Pigulski, K. Shoyama, M.-J. Sun, F. Würthner, J. Am. Chem. Soc. 2022, 144, 5718. 132 X. Yang, M. Brückner, F. Rominger, T. Kirschbaum, M. Mastalerz, Chem 2024, 10, 832. Chapter 3 93 substitution with alkyl side chains leads to an overall 282º contortion of the π- backbone, along the longitudinal axes, inducing chirality (Figure 2.3, twistacenes). Single-crystal X-ray diffraction confirmed the formation of the dimers and revealed that the discrete self-assembly of the twisted monomers is based on multiple weak non-directing interactions involving the side chains, while the π-π interactions are not dominating interactions. Furthermore, the formation of the chiral dimer 3.3 follows the sergeant and soldier principle for the chiral induction. Although the cove regions (in orange) may lead to different helicities along the backbone, the dimer structure is formed between all-M monomers or all-P monomers. Figure 3.2. Solvent dependent van der Waals bilayers leading to chirality induction (3.3) and aggregation induced emission (3.4). As above describe, the formation of van der Waals bilayers has a great impact over the properties. Although the bilayers formation depends mainly on the structure design, the peripheral substitution and the solvent have an important role. In this context, our research group have recently described a Csp3-centered molecular nanographene, tetrahedraphene, that shows aggregation induced emission.133 TD- DFT calculations showed that adding a solvent in which the monomer is insoluble promotes the dimerization leading to bilayer 3.4, which changes the fluorescence color from blue (monomer) to yellow (dimer). This variation may be attributed to the 133 A. Oró, F. Romeo-Gella, J. Perles, J. M. Fernández-García, I. Corral, N. Martín, Angew. Chem. Int. Ed. 2023, 62, e202312314. Chapter 3 94 restriction of the rotational and vibrational motions in the bilayer structure, thus preventing the non-radiative relaxation of the excited state. 3.1.2 Bilayers from Fused Radicals The formation of the so-called pancake bonded nanographenes are currently emerging as a feasible approach to bilayer structures. The control over the stability of open-shell structures (mainly radical or diradical species) lies on preventing their intrinsic trend to undergo σ- or π-dimerization. However, this behavior has proven to be suitable for the controlled formation of persistent bilayer structures. Even though the dimerization of radicals is a reversible process, the predominance of the bilayers can be improved by π-π interactions. Following the previously described strategy to obtain van der Waals bilayer nanographenes (Figure 3.1), Tan, Wang and coworkers prepared the bilayer structure 3.5.134 After reduction reaction with potassium, the structure is transformed into a triply negatively charge radical bilayer 3.53·−, represented in Figure 3.3. The π-dimer radical anion displays a 96-centre-3-electron pancake bond. Interestingly, the open- shell bilayer shows spin frustration, resulting of great interest in magnetic materials physics. It may indicate that upon reduction doping, π-fused nanographenes could present quantum spin liquid properties. Figure 3.3. Bilayers from π- and σ-fused radicals. The well-known σ-dimerization of open-shell molecules has also been explored for the preparation of bilayer structures. In this regard, Stępień and coworkers reported bilayer 3.6 resulting from the formation of a σ-bond between two radicaloid monomers consisting of a hexaazacoronene functionalized with six annulated 134 W. Wang, X.-H. Ma, M. Liu, S. Tang, X. Ding, Y. Zhao, Y.-Z. Tan, M. Kertesz, X. Wang, Angew. Chem. 2023, 135, e202217788. Chapter 3 95 naphthalenemonoimide units, as shown in Figure 3.3.135 The combination of dispersion and covalent forces with the packaging of sterically hindered groups contribute to the high thermodynamic stability of the bilayer dimer. The dissociation process to the open-shell monomers was achieved by irradiation with UV or visible light, providing an opportunity to design magnetically active nanographenes. 3.1.3 Covalently Linked Bilayers The controlled preparation of bilayer and multilayer structures described in the previous sections allows the modulation of the optoelectronic and magnetic properties. However, the structural requirements to obtain persistent bilayers by these approaches −planar π-extended monomers− prevents the induction of structural inherent chirality. As described in Chapter 2.1.1 (Figure 2.3), chirality in nanographenes involves deviation from planarity by the introduction of topological defects. Therefore, an accessible approach is synthesizing bilayer structures, with adjustable optoelectronic and chiroptical properties, in which the overlapping layers are linked by covalent bonds. Our research group was pioneered in describing the first helical bilayer nanographene in 2018.136 The structure of the new molecular nanographene 3.7 consisted of two hexa-peri-hexabenzocoronene moieties annulated to a helicene backbone resulting in a totally conjugated [10]helical bilayer nanographene, Figure 3.4. The formation of the helical feature diverts the structure from planarity inducing chirality and disposing the layers face-to-face at an interlayer distance of 3.6 Å. In the same year Jux and co- workers reported helical nanographene 3.8 consisting of an oxa[7]helicene fused to two HBC layers.137 The main difference between these structures is the length and shape of the helical backbone, which determines the overlapping between the graphitized layers, and the optoelectronic and chiroptical properties. The number of turns of the helicene −and the overlapping between the layers at the ends− depends on the number and geometry of the ortho-fused rings. The presence of five-membered rings opens the inner rim of the helicene decreasing the number of turns (defined as the cycle number of helix). Thus, [10]helical nanographene 3.7 is a partially overlapped structure, with 14 rings involved in the π-π intramolecular interactions, while the length and shape of the helicene in oxa[7]helical nanographene 3.8 prevents the overlapping between the graphitized layers. 135 L. Moshniaha, M. Żyła-Karwowska, P. J. Chmielewski, T. Lis, J. Cybińska, E. Gońka, J. Oschwald, T. Drewello, S. Medina Rivero, J. Casado, M. Stępień, J. Am. Chem. Soc. 2020, 142, 3626. 136 P. J. Evans, J. Ouyang, L. Favereau, J. Crassous, I. Fernández, J. Perles, N. Martín, Angew. Chem. Int. Ed. 2018, 57, 6774. 137 D. Reger, P. Haines, F. W. Heinemann, D. M. Guldi, N. Jux, Angew. Chem. Int. Ed. 2018, 57, 5938. Chapter 3 96 Figure 3.4. Helical bilayer nanographene and π-extended oxa[7]helicene. Moreover, seven-membered rings embedded in the helical feature close the inner rim increasing the number of turns. In this regard, Feng and coworkers reported in 2023 the non-benzenoid [10]helical bilayer nanographene 3.9 (Figure 3.5).138 The presence of two seven-membered rings in the helicene backbone entails the closing of the inner rim leading to the total overlapping between the layers, and a small interlayer distance of 3.24 Å. The in situ registration of UV-vis-NIR spectra upon oxidation, revealed an intervalence charge transfer band in the NIR region for the oxidized species. TD-DFT calculations supported the through-space electronic communication between the layers. In addition, these helical nanographenes are showing highlighting chiroptical absorption (circular dichroism, CD) and emission (circularly polarized luminescence, CPL) properties, determined by the dissymmetry factors of absorption (gabs) and luminescence (glum). For instance, the enantiomers of nanographene 3.9 present a luminescence dissymmetry factor glum of 1.3·10−3. Figure 3.5. π-Extended helical bilayer nanographenes. Also in 2023, Gong et al. reported helical bilayer nanographene 3.10 consisting of a pentadecabenzo[9]helicene core and four fused hexabenzocoronene units (Figure 3.5).139 The rigidity of the structure, due to the π-extension of the helicene, and the large overlapping of the layers (28 benzene rings involved in the π-π interactions) 138 L. Yang, Y.-Y. Ju, M. A. Medel, Y. Fu, H. Komber, E. Dmitrieva, J.-J. Zhang, S. Obermann, A. G. Campaña, J. Ma, X. Feng, Angew. Chem. Int. Ed. 2023, 62, e202216193. 139 Y.-J. Shen, N.-T. Yao, L.-N. Diao, Y. Yang, X.-L. Chen, H.-Y. Gong, Angew. Chem. Int. Ed. 2023, 62, e202300840. Chapter 3 97 affords an interlayer distance of 2.9 Å. The dissymmetry factors for the isolated enantiomers of this nanographene 3.10 are among the highest reported for organic molecules, glum= 4.5·10−2 for the P isomer, and glum= 4.22·10−2 for the M isomer. CPL-active materials140 have attracted considerable interest due to their potential applications in different fields, including photoelectronic devices, data storage, sensing, organic light-emitting diodes, chiroptical switching, etc.141 As previously mentioned, the CPL intensity is described by the dissymmetry factor glum, which represents the ratio of the difference in intensity and the average total luminescence intensity: glum= (IL−IR)/[(1/2)(IL+IR)], IL intensity of left-circularly polarized light, IR intensity of right-circularly polarized light. Theoretically, the dissymmetry factor can be calculated as function of the electric (μ) and magnetic (m) dipole transition moments, glum= 4Re[(μ·m)/(|m|2+|μ|2)]. For organic molecular CPL emitters, the transitions are usually electric dipole allowed and magnetic dipole forbidden. Thus, the magnetic dipole term is much smaller than the electric dipole, and the denominator is dominated by |μ|2, which results in |glum| values in the order of 10−5 to 10−2.142 In addition to the independent study of bilayer molecular nanographenes due to their highlighting chiroptical properties, the last year has witnessed a boom in structure/property relationship studies on bilayer and multilayer nanographenes. Hence, Tan and co-workers reported in 2023 the trilayer nanographenes 3.11 and 3.12 (Figure 3.6), consisting of three HBC moieties connected by [8]helicenes with different topologies.143 As mentioned above, the incorporation of five-membered rings in the helicene causes the inner rim to open, directly related to the overlapping decrease. Therefore, although the number of rings forming the helicenes in both structures is the same, [8]helicenes, the five-membered rings in structure 3.12 cause a decrease in the overlapping between the layers. The determination of the optoelectronic properties showed that three-layer nanographene with higher interlayer overlapping 3.11, has a lower first oxidation potential, stronger bound excitons, and longer photoluminescence lifetimes, thus demonstrating the relationship between the overlapping degree and the optoelectronic properties. 140 Y. Zhang, S. Yu, B. Han, Y. Zhou, X. Zhang, X. Gao, Z. Tang, Matter 2022, 5, 837. 141 Y. Deng, M. Wang, Y. Zhuang, S. Liu, W. Huang, Q. Zhao, Light Sci. Appl. 2021, 10, 76. 142 E. M. Sánchez-Carnerero, A. R. Agarrabeitia, F. Moreno, B. L. Maroto, G. Muller, M. J. Ortiz, S. de la Moya, Chem. Eur. J. 2015, 21, 13488. 143 Y.-Y. Ju, L. Chai, K. Li, J.-F. Xing, X.-H. Ma, Z.-L. Qiu, X.-J. Zhao, J. Zhu, Y.-Z. Tan, J. Am. Chem. Soc. 2023, 145, 2815. Chapter 3 98 Figure 3.6. Helical trilayer nanographenes with overlapping degree depending on the helicenes rings-topology. Further than the comparative study of the properties depending on the overlapping degree, Feng et al. have reported three helical nanographenes, Figure 3.7,144 in which the overlapping degree between contiguous layers remains constant, while the number of layers and the π-extension increase. In this case, the multilayered structures based on covalently connected HBC units forming [7]helicenes, exhibit low overlapping degrees between contiguous layers. However, the global π-π intramolecular interactions increases with the number of layers, attributable to the accumulated interlayer interactions. Higher number of layers enhances the rigidity of the structures, which reduces the interlayer spacing. Thus, the distances are 3.78 Å in bilayer 3.13, 3.26 Å and 3.57 Å in trilayer 3.14, and 3.56 Å, 3.67 Å and 4.20 Å for the contiguous layers in tetralayer 3.15 (from left to right in Figure 3.7). The incremented π-extension at higher number of layers seems to be the structural reason for the properties’ variation. Tetralayer nanographene 3.15 exhibits more red-shifted emission bands, preceded by trilayer 3.14 and bilayer 3.13. This behavior is also observed in the band gap variation, the lowest value corresponds to the tetralayer 3.15. Furthermore, the electrochemical properties show that the lowest oxidation potential corresponds to the most π-extended structure 3.15, which exhibits the strongest donor character. Interestingly, this family of nanographenes present highlighting fluorescence quantum yields, 45% for bilayer 3.13, 74% for trilayer 3.14, and 91% for tetralayer 3.15. This trend in the quantum yields may be related to the number of layers and the consequent rigidity enhancement. The chiroptical properties revealed gradually declined values of the luminescence dissymmetry factors (glum) with the increasing number of layers, 7.9·10−3 for bilayer 3.13, 2.6·10−3 for trilayer 3.14, and 1.5·10−3 for tetralayer 3.15. 144 W. Niu, Y. Fu, Z.-L. Qiu, C. J. Schürmann, S. Obermann, F. Liu, A. A. Popov, H. Komber, J. Ma, X. Feng, J. Am. Chem. Soc. 2023, 145, 26824. Chapter 3 99 Figure 3.7. Helical nanographenes with different number of layers. Also very recently, Feng and coworkers have described the structure/property relationship between three structural isomers of trilayer nanographenes consisting of three HBC units connected by two [7]helicenes.145 The modification of the relative fusion positions between the helicenes and the central benzene ring, affords three different geometries, ortho-fused nanographene 3.14 (Figure 3.7), meta-fused nanographene 3.16, and para-fused nanographene 3.17 (Figure 3.8). Interestingly, the structural modifications entail the amplification of the chiroptical responses, affording high CD and CPL dissymmetric factors; glum values, 8.7·10−3 for M-3.16, and 13.2·10−3 for P-3.17. A 3.2-fold and 4.8-fold amplification compared to the glum value of o-3.14. Figure 3.8. Regioisomeric helical trilayer nanographenes. These recent results are evidence of the growing interest in bilayer and multilayer nanographenes. Just as deviation from planarity by incorporating topological defects entails the induction of chirality, it is feasible to think that enhancing through-space intramolecular π-π interactions represents a new dimension in the design of 145 W. Niu, Y. Fu, Q. Deng, Z.-L. Qiu, F. Liu, A. A. Popov, H. Komber, J. Ma, X. Feng, Angew. Chem. Int. Ed. 2024, 63, e202319874. Chapter 3 100 nanographenes with advanced properties. A leap out of the designing toolbox in which most structures have so far been encompassed, which only considered modifications in the hexagonal network topology and the π-extension. Chapter 3 101 3.2 Objectives In 2018, our research group was pioneer reporting the first helical bilayer nanographene (HBNG). Starting from dibromo[6]helicene, the synthesis afforded a [10]helicene with two HBCs fused at each end.136 The planarity distortion induced by the helical feature, in addition to inducing chirality, dispose the graphitized layers face-to-face, providing a fully conjugated partially overlapped molecular nanographene. The first objective of this chapter is the synthesis and characterization of two helical bilayer nanographenes with different helicene lengths, [9]helicene and [11]helicene. The variation in the number of ortho-fused rings modulates the overlapping degree and the extension of the intramolecular π-π interactions between the layers. The second and main objective of this chapter is the systematic study of the structure/property relationship in this family of bilayer nanographenes with different helicene lengths. For this purpose, the comparative study of the electrochemical and spectroelectrochemical properties, and the photophysical characterization, including Raman analysis, have been carried out. Furthermore, given the chirality of the structures, the absorption and emission chiroptical properties have been evaluated by circular dichroism (CD) and circularly polarized luminescence (CPL). In addition, theoretical calculations have been carried out to support the experimental observations for the structure/property correlations. 102 Chapter 3 103 3.3 Results and Discussion 3.3.1 Synthesis of Helical Bilayer Nanographenes Two chiral bilayer nanographenes have been synthesized from different length helicenes, [5]helicene and [7]helicene. As detailed in Scheme 3.1, the synthesis starts with a Sonogashira coupling between dichloro[5]helicene (18a) or dichloro[7]helicene (18b), and (triisopropylsilyl)acetylene leading to the protected dialkynylhelicenes 19a and 19b, respectively. Subsequently, the one pot deprotection of the TIPS and Sonogashira coupling with 4-tert-butyliodobenzene adds a 4-tert- butylphenyl group to each triple bond affording compounds 20a and 20b. Thereafter, a two-fold Diels-Alder cycloaddition, followed by carbon monoxide extrusion, with cyclone 5 provides the corresponding helicene endowed with two penta(4-tert- butylphenyl)phenyl groups, 21a and 21b. Finally, twelve C−C bonds are formed in the Scholl cyclodehydrogenation with DDQ/TfOH, increasing the number of ortho- fused rings in the helicenes by four, and providing the bilayer molecular nanographenes [9]HBNG and [11]HBNG. Scheme 3.1. Synthesis of helical bilayer nanographenes (HBNGs) with different helicene lengths [9]HBNG, [10]HBNG and [11]HBNG. The structural characterization was carried out by 1H and 13C-NMR, FT-IR and high- resolution mass spectrometry, which allowed the confirmation of all intermediates and final products. The mass spectra of the graphitized structures [9]HBNG (exact mass= 1726.9267) and [11]HBNG (exact mass= 1826.9566) show a loss of 24 units with respect to the molecular weight of the starting compounds 21a (exact mass= 1751.1151) and 21b (exact mass= 1851.1429), respectively, confirming the cyclodehydrogenative formation of 12 bonds. NMR spectra, show the expected Chapter 3 104 signals corresponding to a half of the molecules, revealing the presence of a C2 rotational axis (Appendix 5, spectra). The structures of [9]HBNG and [11]HBNG were solved by single-crystal X-ray diffraction. In both cases, centrosymmetric crystals containing M and P isomers were obtained from the corresponding racemic mixtures, which contrast with the spontaneous independent crystallization of the enantiomers for [10]HBNG. The diverse shapes of the three HBNGs afforded significantly different packing, Figure 3.9a. As expected, the helicene length determines the bilayer character of the structures, namely the overlapping degree between the facing HBC units. As represented in Figure 3.9b, the number of rings involved in the π-π interactions (represented with dashed lines) between the graphitized layers changes drastically from [9]HBNG to [11]HBNG. Structure [9]HBNG is a totally overlapped bilayer with 26 rings out of 29 involved in the π-π intramolecular interactions, arranging the layers at an average distance of 3.63 Å. The increased number of turns induced by the longer helicene length in [10]HBNG causes the displacement of the graphitized layers, thereby leading to a partially overlapped bilayer in which 14 rings participate in π-π interactions. In the case of [11]HBNG, the extension of the helicene prevents the face-to-face placement of the HBC layers, resulting in a shifted bilayer in which the 10 rings involved in the π-π interactions are part of the helicene backbone. This contrasts with [9]HBNG structure in which the π-π interacting rings are mainly located in the HBC units. Figure 3.9.(a) Packing of the structures [9]HBNG, [10]HBNG and [11]HBNG. (b) Lateral views of molecular structures of [9]HBNG, [10]HBNG and [11]HBNG, with dashed lines are represented the rings involved in the intramolecular π-π interactions showing the overlapping degree variations, tert-butyl groups and hydrogens have been removed for clarity. Higher overlapping [9]HBNG [10]HBNG [11]HBNG (a) (b) Chapter 3 105 3.3.2 Electrochemical and Spectroelectrochemical Properties The electrochemical properties of [9]HBNG, [10]HBNG, and [11]HBNG were evaluated by cyclic voltammetry in a 0.1 M 4:1 toluene/acetonitrile solution of tetrabutylammonium hexafluorophosphate at room temperature, using Ag/AgNO3 as reference electrode, glassy carbon as working electrode, and a platinum wire as counter electrode. In Table 3.1 are summarized the oxidation and reduction potentials vs. Fc/Fc+, including the data corresponding to tBu-HBC for comparison. As shown in Figure 3.10, all three bilayer nanographenes present two quasi-reversible oxidation waves and two quasi-reversible reduction waves. The values for the first oxidation potentials follow the order [9]HBNG (0.35 V) < [10]HBNG (0.46 V) < [11]HBNG (0.52 V), showing a significantly stronger donor character than tBu-HBC (0.75 V) due to the larger π-extension of HBNGs. However, the donor character behavior of HBNGs is opposite to that expected according to the π-extension of the bilayer nanographenes. The strongest electron donor is the least π-extended structure, [9]HBNG, and the weaker electron donor is the largest π-extended structure, [11]HBNG. Figure 3.10. Cyclic voltammograms of HBNGs, [9]HBNG, [10]HBNG, and [11]HBNG vs. Fc/Fc+ in a 1 M solution of tetrabutylammonium hexafluorophosphate in toluene/acetonitrile 4:1. The oxidation potential variations suggest that the larger the π-π overlapping (Figure 3.9), the lower the oxidation potential values. This observation could be accounted for by the fact that the radical cation and dication species (Eox 1 and Eox 2, respectively) might be stabilized through the interaction between the two graphitized layers. This observation agrees with the density functional calculation of the HOMO, which is mainly located in the HBC units, and their energies vary in line with the measured oxidation potentials (Figure 3.11). Chapter 3 106 Table 3.1. Half-wave redox potentials of tBu-HBC, [9]HBNG, [10]HBNG and [11]HBNG vs Fc/Fc+. Compound E1 1/2ox (V) E2 1/2ox (V) E1 1/2red (V) E2 1/2red (V) tBu-HBC 0.75 – –2.24 –2.40 [9]HBNG 0.35 0.59 –2.18 –2.46 [10]HBNG 0.46 0.67 –2.23 –2.55 [11]HBNG 0.52 0.69 –2.22 –2.46 In terms of the reduction potentials, the variation is less prominent, being [9]HBNG (−2.18 V), [10]HBNG (−2.23 V), and [11]HBNG (−2.22 V). According to DFT calculations the LUMO of the three structures is mainly located in the helicene moiety (Figure 3.11). In this sense, the scarce bond conjugation along the helicenes may explain the minor differences observed in the reduction potentials. Figure 3.11. Density functional theory calculations of the HOMO and LUMO topologies and energies. To determine whether the radical cation and dication species are indeed stabilized between the layers, and its relation to the stronger donor character of [9]HBNG (totally overlapped bilayer), UV-vis-NIR spectroelectrochemical measurements were carried out. Using platinum as working electrode, a platinum wire as counter electrode, and a silver wire as the reference electrode, in a 0.1 M solution of tetrabutylammonium hexafluorophosphate in dichloromethane. −0.694 −5.762 −5.839 −0.699 −5.851 [9]HBNG [10]HBNG [11]HBNG B 3 L Y P /6 -3 1 G ** e n e rg y LUMO HOMO −0.684 Chapter 3 107 As shown in Figure 3.12b (blue line), the first oxidation of [9]HBNG involves the appearance of two absorption bands at 934 and 540 nm, comparable with the bands of tBu-HBC at 797 and 550 nm (Figure 3.12a, blue line), upon formation of the radical cation. The redshift of the absorption bands from 797 to 934 nm agrees with the lower oxidation potential of [9]HBNG relative to tBu-HBC. Thus, the significant decrease of the oxidation potential observed for [9]HBNG (vs. tBu-HBC) suggests a double effect: (i) the involvement of the helicene moiety in the oxidation and (ii) considering the redshift of the absorption bands, an interlayer charge delocalization, which eventually contributes to decrease the oxidation potential. Figure 3.12. UV-Vis-NIR electronic absorption spectra obtained upon electrochemical oxidation of a) tBu-HBC, b) [9]HBNG. Black lines correspond to neutral species, blue lines to the first oxidized species and red lines to the second oxidized species. Additionally, the most intense absorption band at 373 nm of [9]HBNG in the neutral state (Figure 3.12b, black line) progressively shifts to the blue with the first, 362 nm (blue line), and the second oxidations, 333 nm (red line). Since this intense sharp band in the neutral state is composed of transitions emerging from the helicene and the HBC moieties, the described shifting upon oxidation supports that, the charge is jointly stabilized by the two structural units. 3.3.3 Photophysical Properties The spectroscopic properties of the bilayer nanographenes were evaluated by UV-vis absorption (summarized in Table 3.2) and emission measurements (summarized in Table 3.3). Bilayer nanographenes [9]HBNG, [10]HBNG and [11]HBNG display non-structured and broad bands in the absorption spectra (Figure 3.13, solid lines), which contrast to the structured absorption bands corresponding to tBu-HBC. HBNGs show slightly bathochromically sifted maxima in the 373-377 nm interval, compared to the tBu-HBC maximum at 360 nm. Interestingly, bilayer nanographenes show allowed electronic transitions in the visible region above 400 nm (absorption coefficients >104 M−1·cm−1). (a) (b)tBu-HBC [9]HBNG 540 934 362 333 Chapter 3 108 Table 3.2. Data from absorption spectra of the HBNGs and tBu-HBC. Compound λabs max /nm (ε /M−1·cm−1) tBu-HBC 344 (63,750); 360 (141,700); 390 (46,350); 439 (1,600); 441 (1,650); 445 (1,800) [9]HBNG 349 (94,600); 373 (136,000); 392 (86,300); 435 (20,600); 502 (6,200); 542 (4,700) [10]HBNG 360 (98,600); 377 (103,900); 415 (38,800); 446 (18,700); 462 (14,400); 494 (8,700) [11]HBNG 339 (70,800); 377 (125,000); 415 (58,200); 444 (26,100); 491 (7,300) The emission spectra of the three bilayer nanographenes show broad red-shifted non- structured bands with maxima at 575 nm ([9]HBNG), 543 nm ([10]HBNG), and 528 nm ([11]HBNG), while the rigid and flat monolayer nanographene tBu-HBC shows a well-structured spectrum with a maximum at 484 nm. Interestingly, the lower π- extended and totally overlapped structure (with larger π-π interactions, Figure 3.9) [9]HBNG, shows the most bathochromically shifted fluorescence, being in agreement with the electrochemical properties previously described ([9]HBNG shows the strongest donor character). As illustrated in Figure 3.13 (shaded curves), the emission spectra of HBNGs show two main features: (i) in shape, they resemble the emission spectra of helicenes with one molecular band disclosing vibronic resolution and (ii) the pronounced broadening is typical of excimer-type emissions. The most red-shifted excimer-like band of [9]HBNG, agrees with larger π-π interactions resulting from the total overlapping of this bilayer. Figure 3.13. Absorption (solid lines) and emission (shaded curves) spectra of tBu-HBC, [9]HBNG, [10]HBNG and [11]HBNG. Picture in the right show samples of the three bilayer nanographenes in CHCl3 under 365 nm irradiation. Chapter 3 109 The fluorescence quantum yields were determined by comparison with the emission spectra of riboflavin in ethanol (Φem= 0.3±0.03). The variations in the obtained fluorescence quantum yields, Φem= 0.22 for [9]HBNG > Φem= 0.10 for [10]HBNG ~ Φem= 0.11 for [11]HBNG, do not follow the expected order considering the band gaps (estimated from the intersection between the normalized excitation and emission spectra), E0-0[9]HBNG= 2.35±0.09 eV < E0-0[10]HBNG= 2.49±0.09 eV < and E0-0 [11]HBNG= 2.58±0.09 eV. However, this can be partially attributed to the different contributions of excimer-like emission in HBNGs. Hence, the structure with the lower band gap, totally overlapped [9]HBNG, presents the highest fluorescence quantum yield due to the more accentuated excimer-like emission. Table 3.3. Data from excitation and fluorescence spectra of the HBNGs and tBu-HBC. Compound λem max /nm (shoulder) E0-0 /eV Φem tBu-HBC 445, 454, 464, 475, 484, 493, 518, 528, (537), 556, 567 2.65±0.13 0.02±0.01 [9]HBNG 575 (611) 2.35±0.09 0.22±0.02 [10]HBNG 543 (567) 2.49±0.09 0.10±0.01 [11]HBNG 528 (551) 2.58±0.09 0.11±0.01 The fluorescence decays of HBNGs are shown in Figure 3.14. While [9]HBNG and [10]HBNG required multiexponential fittings of the decay function, [11]HBNG exhibited a monoexponential behavior. The calculation of the discrete lifetime components (see Appendix 3, photophysical details) provided the intensity-weighted average lifetimes, 13.4 ns for [9]HBNG, 12.2 ns for [10]HBNG, and 8.7 ns for [11]HBNG. Figure 3.14. Fluorescence decay profiles of [9]HBNG (left), [10]HBNG (middle) and [11]HBNG (right) in chloroform, obtained at different emission detection wavelengths with excitation at 405 nm. Instrument response functions (IRF) are shown in red. From the amplitude-weighted fluorescence average lifetimes, and fluorescence quantum yields, the radiative (kr) and non-radiative (knr) deactivation rate constants of the respective singlet excited states were calculated. All three structures present similar kr (ca 1−2·107 s−1). However, there is a significant difference between non- 0 50 100 150 1E-4 0.001 0.01 0.1 1 N o rm a liz e d i n te n s it y t[ns] HB-11 IRF 530 nm 560 nm [11]HBNG IRF 530 nm 560 nm 0 50 100 150 1E-4 0.001 0.01 0.1 1 N o rm a liz e d i n te n s it y t[ns] HB-10 IRF 507 nm 540 nm 570 nm 630 nm [10]HBNG IRF 507 nm 540 nm 570 nm 630 nm 0 50 100 150 1E-4 0.001 0.01 0.1 1 N o rm a liz e d i n te n s it y t[ns] HB-9 IRF 512 nm 578 nm 620 nm [9]HBNG IRF 512 nm 578 nm 620 nm Chapter 3 110 radiative deactivation constants (knr), being of 7.0·107 s−1 for totally overlapped [9]HBNG, 1.5·108 s−1 for partially overlapped [10]HBNG, and 1.0·108 s−1 for the shifted structure [11]HBNG, which means that the layer overlapping, and subsequent stronger π-π intramolecular stacking, hinders the structural mobility favoring the radiative deactivation. The greater the overlap, the higher the emission quantum yields and the longer the fluorescence lifetimes. The three bilayer nanographenes were characterized by Raman spectroscopy in solid state using different excitation laser wavelengths to avoid the adverse effect of fluorescence. For [10]HBNG and [11]HBNG, Raman spectra were obtained at 1064 nm. For [9]HBNG they were obtained by excitation in the UV region with the 325 nm laser Raman line, since the intense fluorescence and reabsorption in the NIR region prevented obtaining quality Raman spectra at 1064 nm. As shown in Figure 3.15, all Raman spectra present two main groups of bands at 1600 and 1350-1300 cm−1, which correspond to the Raman G and D bands typical of vibrational Raman spectra of graphene (see Appendix 3, Figure AP3.2, theoretical Raman spectra and assignments of the most representative Raman bands in terms of vibrational normal modes). In particular, the Raman spectrum of [10]HBNG represents the molecular version of the full Raman spectrum of graphene and multilayer graphene (Figure 3.15, left). The face-to-face packing of the graphitized layers leads to the activation and appearance of bands associated with the double and triple wavenumbers of the fundamental phonon Raman vibrations. The appearance of these multiphonon bands at 2779 and 2711 cm−1 (i.e., 2D and 2G) is an intrinsic feature observed when switching from graphene to multilayer graphene. Figure 3.15. Left, solid-state Raman spectrum of [10]HBNG. Right, solid-state Raman spectra of tBu-HBC (a), [9]HBNG (b), [10]HBNG (c), and [11]HBNG (d). The Raman spectrum of [10]HBNG shows the highest similarity to that of tBu-HBC in the region of the G Raman band at 1613 cm−1, whereas the G band of [9]HBNG at 1630 cm−1 shows the highest difference from tBu-HBC. This variation from 1613 (tBu- HBC) to 1630 cm−1 ([9]HBNG) may be related to the vibrational mixing of the C−C bond stretching of the HBC moiety with that of the helicene (Appendix 3, Figure AP3.3). 3000 2700 2400 2100 1800 1500 1200 900 600 1379 1341 1361 N o rm . R a m a n I n t. Raman Shift (cm -1 ) 2711 2779 1800 1700 1600 1500 1400 1300 1200 1100 1000 a d c b1370 13261355 1605 13411361 1613 13671630 13231601 N o rm . R a m a n I n t. Raman Shift (cm -1 ) 1613 Chapter 3 111 3.3.4 Chiroptical Properties of HBNGs The racemic mixtures of HBNGs were resolved by means of semipreparative chiral HPLC (Chiralpak IE column) using heptane and 1% isopropyl alcohol in toluene as mobile phase. All enantioenriched samples showed remarkable enantiomeric excess (between 98-99 ee), except for (+)-[10]HBNG that was isolated with a 73% ee (Appendix 3). Figure 3.16. CD spectra of the enantioenriched samples of HBNGs, including the gabs values. Scaled by a factor of 1.4 for [10]HBNG considering the purity. Chapter 3 112 The circular dichroism of the six enantioenriched samples showed specular image signals for each pair of enantiomers. As illustrated in Figure 3.16, the CD spectra of [10]HBNG and [11]HBNG present similar Cotton effects in the visible region, with distinguishable contributions corresponding to the helicene backbone and the HBC at 380 and 450 nm, respectively. Yet, the CD spectra of [9]HBNG show three regions with prominent cotton effects in the visible range, the one observed for the longer helical bilayers, at 380 nm, and two more at 400 nm and 450 nm, evidencing the mixed contributions of the helicene and the HBC, and suggesting the involvement of the overlapping in the chiroptical properties. The circularly polarized luminescence (CPL) was also recorded for the six enantioenriched chiral bilayer nanographenes, Figure 3.17. The three structures exhibit intense CPL signals with remarkable asymmetry factors (i.e., glum= 2(IL−IR)/(IL+IR), IL and IR being the left- and right-handed luminescent emissions, respectively). Figure 3.17. Circularly polarized luminescent (CPL) and normalized emission (PL) spectra of (+)- and (−)-HBNGs. The |glum| values, as shown in Table 3.4, range from 1.0·10−2 to 3.6·10−2, and are of the same order as the circular dichroism dissymmetry factors (gabs), reflecting small structural and electronic reorganization of the excited state prior to the emission process. Interestingly, the much higher values were obtained for [9]HBNG (|glum|= 3.610−2 at 580 nm), being remarkable for structures formed only by carbon and hydrogen. As proposed by Zinna and co-workers,146 CPL brightness should be considered as a unified measure of circularly polarized emission properties. This value is calculated from the absorption coefficient, the CPL dissymmetry factor, and the quantum yield (BCPL= ε·Φem·glum/2). Thus, the molecular nanographenes described 146 L. Arrico, L. Di Bari, F. Zinna, Chem. Eur. J. 2021, 27, 2920 Wavelength (nm) Δ I / 1 0 − 2 −4 −2 0 2 4 N o rm a liz e d P L 1 0 (−)-[9]HBNG (−)-[10]HBNG (+)-[9]HBNG (−)-[11]HBNG (+)-[10]HBNG (+)-[11]HBNG 450 500 550 600 650 700 750 800 Chapter 3 113 have the following values, BCPL[9]HBNG= 81, BCPL[10]HBNG= 21, and BCPL[11]HBNG= 28. Table 3.4. Experimental dissymmetry factors from circular dichroism (gabs) and circularly polarized luminescence (glum). Compound gabs /10−2 (λ, nm) glum /10−2 (λ, nm) (+)-[9]HBNG +3.6 (540) +3.6 (580) (−)-[9]HBNG −2.8 (540) −3.6 (580) (+)-[10]HBNG +1.4 (502) +1.1 (540) (−)-[10]HBNG −1.6 (502) −1.0 (540) (+)-[11]HBNG +1.0 (490) +0.8 (535) (−)-[11]HBNG −1.0 (490) −0.9 (535) The experimental glum values agree with the theoretically calculated (TD-DFT) from the quantum chemical quantities, the electric transition dipole moment (ETDM, μ) and the magnetic transition dipole moment (MTDM, m) according to the equations; R= |μ||m|cosθ and gabs= (4|m|/|μ|)cosθ. Adopting the accepted approach of discussing the experimental glum compared with the computationally available gabs when emission and absorption correspond to the same lowest energy electron transition, the equation can be also considered as glum= (4|m|/|μ|)cosθ. The TD-DFT calculations nicely predict the experimental order of magnitude and variation of the glum for the three compounds, as shown in Figure 3.18. Chapter 3 114 Figure 3.18. Direction, moduli and values of the ETDM (blue arrow) and MTDM (red arrows), angles between the vectors (θ), and calculated glum values for [9]HBNG, [10]HBNG and [11]HBNG, calculated at the optimized CAM-B3LYP/3-21G* geometry. The moduli of the ETDM and MTDM transition vectors are higher for [9]HBNG, and display the smallest angle (θ= 3.4º, smallest angle entails largest cosine); thus, resulting in higher dissymmetry factor value for [9]HBNG. The structural design of [9]HBNG results optimal for increasing the gabs and glum. The balanced π-extension of the chromophore between the planar HBC (promoting large ETDM) and the helicene (promoting large MTDM) is an appealing alternative to obtain good emission and chiroptical properties. ETDM |μ|=1.0160·10−18 esu·cm MTDM |m|=2.6270·10−21 erg/Gauss θ=3.4º glum_calc=1.03·10−2 ETDM |μ|=0.9453·10−18 esu·cm MTDM |m|=1.4014·10−21 erg/Gauss θ=17.78º glum_calc=0.56·10−2 ETDM |μ|=0.8385·10−18 esu·cm MTDM |m|=1.7485 ·10−21 erg/Gauss θ=63.51º glum_calc= 0.37·10−2 [9]HBNG [10]HBNG [11]HBNG Chapter 3 115 3.4 Conclusions To conclude, the synthesis and characterization of two new HBNGs, [9]HBNG and [11]HBNG, have been carried out following a methodology similar to that described for [10]HBNG.136 The structures have been unequivocally characterized by single- crystal X-ray diffraction, which has allowed the understanding of the unique bilayer arrangement of the layers. The great variations in the overlapping are clearly shown by the number of rings involved in the π-π interactions, 26 for totally overlapped bilayer [9]HBNG, 14 for partially overlapped bilayer [10]HBNG, and 10 for the shifted nanographene [11]HBNG. The systematic and comparative study of the properties demonstrate the strong influence of the overlapping degree on the optoelectronic properties. Cyclic voltammetry measurements unexpectedly showed that the strongest electron donor character corresponded to the less π-extended structure [9]HBNG. The in situ registration of UV-vis-NIR spectra upon oxidation (spectroelectrochemical measurements) showed a red shifting of the bands at higher wavelengths, in the absorption spectrum corresponding to the radical cation of [9]HBNG. Thereby revealing a charge delocalization between the two HBC layers contributing to a lower oxidation potential, i. e. mixed valence band effect. Actually, this intramolecular through-space communication supports the differences found in the emission spectra of the three bilayers, the most red-shifted maximum corresponds to [9]HBNG, followed by [10]HBNG and [11]HBNG. Additionally, the total overlapping of the layers is related to the greater fluorescence quantum yield of [9]HBNG, the rigidity enhanced by the strong π-π interactions favors the relative radiative deactivation. Interestingly, the dissymmetry factors gabs and glum determining the intensity of the chiral absorption (CD) and emission (CPL) properties, show greater values for the totally overlapped structure [9]HBNG, with values as high as g= 3.6·10−2. These values are among the highest reported in literature for carbon and hydrogen compounds. The experimental findings about this family of HBNGs, demonstrate the influences of the unique intramolecular π-π interactions between the covalently connected HBC layers over the properties. Chapter 4 Chapter 4 119 4 Chapter 4 4.1 Introduction to Stereochemical Control in Nanographenes Chirality is a geometrical property, whereby a structure is not superimposable with its mirror image. As seen in Chapter 3, it gives rise to properties of high interest in advanced materials. However, the determination and application of these properties may be challenging since, as seen in Chapter 2, the retention of properties with time depends on the energy of the isomerization barriers. Moreover, it is necessary to isolate the enantiomers to determine these properties associated to enantioenriched or enantiopure samples. Most chiral nanographenes syntheses described in the literature lead to racemic mixtures, which involves the subsequent separation of the enantiomers by chiral HPLC. These separations are usually complex, time consuming and expensive, as well as being a limitation to obtain enantiopure samples in quantities of more than a few milligrams. 4.1.1 Enantioselective Synthesis of Nanographenes Alternatively, to avoid the separation process, the enantioselective synthesis of nanographenes leading to the controlled formation of enantiomers is very appealing and challenging. The enantioselective, or asymmetric, synthesis aims at the preferential formation of one enantiomer over another during a chemical reaction. For this purpose, a chiral auxiliar component of defined configuration interacts with a prochiral molecule generating diastereotopic reactive positions. This difference between the reactive positions causes the reaction to evolve preferentially through the one with the lowest energy. The configuration of the chiral auxiliary determines the preferential interaction with the prochiral molecule, and therefore the enantioselectivity of the process. In this context, our research group has set a milestone reporting the first enantioselective synthesis of a molecular nanographene.147 As shown in Scheme 4.1, the enantioselective synthesis starts with previously prepared indandione 4.1, containing a polyarene (PAB: pentaaryl benzene). The enantioselective reduction of the carbonyl groups, performed in the presence of the S-(4- methoxy)phenyloxazaborilidine 4.2 and borane, affords the trans-diol S,S-4.3 (enantiomeric excess, e.e.>97%) and meso-4.3 (cis compound) in 70:30 diastereomeric ratio. Then, the enantioenriched diol S,S-4.3 undergoes a stereoselective intramolecular Friedel-Crafts reaction in which the ortho-position at five-bonds distance, attacks the asymmetric carbon with inversion of the configuration (via SN2), affording S,S,Sa-4.4 with an e.e.> 97%. The last step is the Scholl reaction, which eventually leads to the graphitized chiral nanographene S,S,Sa,M,M-4.5 with an enantiomeric excess of 97%. The enantiospecific performance of the final Scholl reaction to form M-[6]helicenes allows the retention of the chiral information. 147 M. Buendía, J. M. Fernández-García, J. Perles, S. Filippone, N. Martín, Nat. Synth. 2024, 3, 545. Chapter 4 120 Moreover, if the first enantioselective reduction of 4.1 is performed in the presence of the R enantiomer of oxazaborilidine 4.2, nanographene R,R,Ra,P,P-4.5 is obtained with a 98% enantiomeric excess. Scheme 4.1. Enantioselective synthesis of a two-fold inherently chiral nanographene. Therefore, the success achieved with this synthesis proves a high stereochemical control for the enantioselective preparation of sophisticated carbon nanostructures. From the first enantioselective reduction step, a chiral asymmetric carbon (point chirality) is generated, subsequently this chirality is transferred to enantioselectively obtain a spirocycle (axial chirality), whose configuration entails the enantiospecific graphitization reaction in which two helicenes are formed (helical chirality). It is worth mentioning that this last step represents the first enantioselective Scholl reaction reported in the literature since its first report in 1910. 4.1.2 Chirality Transference: from Atropisomerism to Helical Chirality Alternatively, axially chiral atropisomers have been frequently considered a powerful chiral auxiliar to control the helical configuration of helicenes and heterohelicenes. The first example was reported in 1972 by Both et al. As represented in Scheme 4.2, Chapter 4 121 the optically pure biphosphonium periodate S-binaphthyl 4.6 with lithium ethoxide afforded optically pure P-[5]helicene (4.7).148 Scheme 4.2. First synthesis of optically pure P-[5]helicene through chirality transference from an optically pure binaphthyl derivative. In 1994, Hanus and coworkers demonstrated the effective intramolecular chirality transference from axial chirality to point chirality, and eventually to helical chirality, Scheme 4.3. The first key step is the chirality transference during the Stevens rearrangement. The enantiopure dinaphthyl azepinium salt S-4.8, under basic conditions affords a single enantiomer R,3R-4.9. The obtained compound R,3R-4.9 in presence of m-chloroperbenzoic acid undergoes a Cope elimination providing the enantiomerically pure P-[5]helicene 4.10, which was also obtained from the enantiomerically pure derivative S-4.8 in presence of butyl lithium.149 Scheme 4.3. Synthesis of optically pure [5]helicene from intramolecular chirality transference. The synthesis of optically pure sulfur-containing heterohelicenes 4.11 was successfully achieved by Osuga et al. in 1997. The optically pure dialdehyde 4.12 was obtained through chemical resolution of the precursors. Thus, the McMurry coupling reaction from compound 4.12 yields the heterohelicene P-4.11 with axial to helical chirality transference, Scheme 4.4.150 148 H. Jürgen Bestmann, W. Both, Angew. Chem. Int. Ed. Engl. 1972, 11, 296. 149 I. G. Stará, I. Starý, M. Tichf, J. Závada, V. Hanus, J. Am. Chem. Soc. 1994, 116, 5084. 150 K. Tanaka, H. Suzuki, H. Osuga, J. Org. Chem. 1997, 62, 4465. Chapter 4 122 Scheme 4.4. Axial to helical chirality transference during McMurry coupling. In 2005 Nozaki and coworkers reported the synthesis of the azahelicene P-4.13 from the enantiomerically pure bisphenantryl derivative S-4.14, obtained by chemical resolution. The stereospecific double N-arylation leads to the aza[7]helicene with optical purity retention, Scheme 4.5.151 Scheme 4.5. Stereospecific double N-arylation to afford an optically pure aza[7]helicene. Alternatively, Collins and coworkers reported in 2008, the first application of olefin metathesis to obtain helicene molecules, Scheme 4.6. They described a kinetic resolution by means of asymmetric ring-closing metathesis from the bisphenanthrene derivative 4.15 in presence of a chiral ruthenium catalyst to afford [7]helicene (4.16) with a maximum 80% e.e.152 Therefore, in this case, the catalyst is the responsible species for the induction of chirality. Scheme 4.6. Kinetic resolution and asymmetric metathesis for the synthesis of carbohelicenes. Starting from enantiopure commercially available R-BINOL, Hu and coworkers reported the diastereoselective photochemical synthesis of BINOL-fused helicenes, 151 K. Nakano, Y. Hidehira, K. Takahashi, T. Hiyama, K. Nozaki, Angew. Chem. Int. Ed. 2005, 44, 7136. 152 A. Grandbois, S. K. Collins, Chem. Eur. J. 2008, 14, 9323. Chapter 4 123 Scheme 4.7. The photocyclization of R,R-4.17 provides the [7]helicenes mixture M,R,R-4.18 and P,R,R-4.18 in 2.5:1 ratio, respectively.153 Scheme 4.7. Stereoselective photochemical synthesis of helicenes. In the following Scheme 4.8, the synthesis of the enantioenriched [5]helicene 4.19 with chirality transference from the enantioenriched atropisomer 4.20, is represented. As already described in previous examples, the optical purity of the starting material with axial chirality determines the optical purity of the helical feature. Therefore, the kinetic, chemical or HPLC resolution of the atropisomers is usually required. In this case, Ackermann and coworkers reported an asymmetric metalla-electrocatalyzed C−H activation for the preparation of the enantioenriched biaryl derivative 4.20.154 Scheme 4.8. Asymmetric metalla-electrocatalysis for the preparation of an atropisomers and chirality transference to a [5]helicene derivative. Yashima et al. have recently reported the simultaneous enantiopure synthesis of the [7]helicene 4.21 and the non-benzenoid[6]helicene 4.22, with opposite helicity, Scheme 4.9. The enantiomerically pure S,S-5,6-di(1-naphthyl)benzene derivative 4.23 153 S. Chen, Z. Ge, Q. Jia, K.-P. Wang, L.-H. Gan, Z.-Q. Hu, Chem. Asian J. 2019, 14, 1462. 154 U. Dhawa, C. Tian, T. Wdowik, J. C. A. Oliveira, J. Hao, L. Ackermann, Angew. Chem. Int. Ed. 2020, 59, 13451. Chapter 4 124 (isolated by chiral HPLC) in presence of trifluoroacetic acid undergoes two stepwise intramolecular alkyne annulations. The helical handedness of the resulting helicenes is determined by the double axial precursor. 155 Scheme 4.9. Enantiopure synthesis of [7]helicene and non-benzenoid[6]helicene derivatives. These examples representatively illustrate the wide variety of reactions that can be carried out with axial-to-helical chirality transference. The suitable design of the precursor, as well as the methodology to be followed, are key to obtain efficient enantiomeric excesses. Typically, the enantiopurity of the axially chiral precursor determines the enantiopurity of the resulting helical feature. In the common case where the precursor presents two enantiomers, chiral HPLC can be used for its isolation. However, an affordable and easy to scale-up methodology is the chemical resolution to obtain diastereomers that can be isolated by a common silica gel chromatography column. 155 T. Ikai, K. Oki, S. Yamakawa, E. Yashima, Angew. Chem. Int. Ed. 2023, 62, e202301836. Chapter 4 125 4.2 Objectives The main objective of this chapter is the enantiospecific synthesis of helical bilayer nanographene. The synthetic design for this purpose consists on the chirality transference from the axial chirality of BINOL-based polyarene derivatives to the helical chirality characteristic of helical bilayer nanographenes (HBNGs, Chapter 3). The strategy is based on the chemical resolution of the polyarene enantiomers with axial chirality for the subsequent enantiospecific Scholl reaction. This easily scalable methodology starts from simple and commercially affordable building blocks, which allows the simplification of the synthetic route while avoiding the costly HPLC separation of enantiomers. Once the synthetic objective was fulfilled, the optoelectronic and chiroptical properties were determined. The resulting structure consists of an oxa[9]helical bilayer nanographene. The detailed analysis of the nanographene by single-crystal X- ray diffraction showed the great overlapping of the layers, which is strongly related to the variation of the optoelectronic properties as described in the previous Chapter 3. However, the electron donor character of the oxygen atom embedded in the helicene backbone, modulates the optoelectronic properties, specifically those related to the electron-acceptor character of the nanographene. 126 Chapter 4 127 4.3 Results and Discussion: oxa-Helical Bilayer Nanographenes 4.3.1 Enantiospecific Synthesis of oxa-Helical Bilayer Nanographenes In this chapter, the synthesis and characterization of oxa[9]helical bilayer nanographene, oxa[9]HBNG, is described. In contrast to the synthesis of [9]HBNG described in Chapter 3.3.1, the synthetic methodology starts from 7-bromo-2-naphtol instead of requiring the previous preparation of dichloro[5]helicene. Thus, as detailed in Scheme 4.1, the first step is an oxidative homocoupling of 7-bromo-2-naphtol (22), mediated by a Cu-TMDA catalyst, affording 7,7'-dibromo-1,1'-binaphthyl-2,2'-diol (23). Then, a two-fold Sonogashira coupling is performed between the 7,7′- dibromobinol 23 and two equivalents of 4-(tert-butyl)phenylacetylene (3), providing 7,7'-bis[(4-(tert-butyl)phenyl)ethynyl]-1,1'-binaphthalyl-2,2'-diol (24). The resulting alkynyl containing BINOL was assembled with the cyclopentadienone 5 to perform a double Diels-Alder cycloaddition, after in situ carbon monoxide extrusion, the BINOL core endowed with two penta(4-tert-butylphenyl)benzene groups is obtained with remarkably good yield, 93%. Finally, the Scholl cyclodehydrogenation is performed with DDQ and triflic acid. During this last step, twelve carbon-carbon bonds are formed providing the graphitization. In addition, the cyclodehydration of the phenol groups yields a carbon-oxygen bond formation affording a furane-embedded helicene backbone, and thus, the helical bilayer nanographene oxa[9]HBNG. Scheme 4.1. Synthesis of racemic oxa[9]HBNG from 7-bromo-2-naphtol. The building block 1,1'-binaphthyl-2,2'-diol, commonly known as BINOL, is a widely used core in synthesis.156 The reactive phenol groups and the axial chirality are the 156 J. M. Brunel, Chem. Rev. 2005, 105, 857. Chapter 4 128 perfect combination for a wide variety of applications. Thus, it has been broadly used in asymmetric catalysis, for the design of sophisticated chiral ligands for catalysts,157,158 but also in materials chemistry, for instance, as a chiral feature linked to photoactive BODIPY dyes,159 and [6]cycloparaphenylene,160 giving rise to remarkable chiroptical properties. Furthermore, Mateo-Alonso et al. have recently reported the use of 1,1′-binaphthyl-2,2′-diamine for the asymmetric chirality induction in a twistacene nanoribbon.161 The covalent binding of two configurationally stable biphenyls at the ends of the nanoribbon induce the twisting of the structure with conformational discrimination. The chirality of 1,1′-binaphthyl derivatives arises from the restricted rotation of the naphthyl units around the single bond. The steric hindrance between the hydrogens in the positions 8,8′ entails a 23 kcal·mol−1 isomerization barrier in 1,1′-binaphthyl. Further functionalization in the 2,2′ positions to obtain 1,1′-binaphthyl-2,2′-diol (BINOL) increases the isomerization barrier up to 38 kcal·mol−1, providing configurational stability at room temperature.162 The incorporation of a BINOL core in the structure, as shown in Scheme 4.1, provides great versatility considering the configurational stability of the isomers. In addition, its reactivity to form furan embedded oxa-helicenes163 provides the possibility of modulating, and improving, the optoelectronic and chiroptical properties, as demonstrated by different studies of heterohelicenes.164,165,166 The gram scale synthesis of helical bilayer nanographenes (HBNGs) is still limited due to several step synthetic routes. As detailed in Chapter 3, these molecular nanographenes have proven remarkable chiroptical properties, however, their application as CD or CPL active materials requires the expensive separation of enantiomers by chiral HPLC. Herein, the easily scalable enantiospecific synthesis of oxa[9]HBNG from low cost commercially available starting materials, is described. As detailed in Scheme 4.2, the enantiopure preparation of oxa[9]HBNG starts with the functionalization of the polyarene 25 with 1R-10-camphorsulphonyl chloride, affording two diastereomers, R,M-26a and R,P-26b, easily isolable by silica gel column chromatography. 157 J. Aydin, K. S. Kumar, M. J. Sayah, O. A. Wallner, K. J. Szabó, J. Org. Chem. 2007, 72, 4689 158 R. Kshatriya, ACS Omega 2023, 8, 17381. 159 J. Jiménez, R. Prieto-Montero, S. Serrano, P. Stachelek, E. Rebollar, B. L. Maroto, F. Moreno, V. Martinez-Martinez, R. Pal, I. García-Moreno, S. de la Moya, Chem. Commun., 2022, 58, 6385. 160 K. Sato, M. Hasegawa, Y. Nojima, N. Hara, T. Nishiuchi, Y. Imai, Y. Mazaki, Chem. Eur. J. 2021, 27,1323. 161 R. K. Dubey, M. Melle-Franco, A. Mateo-Alonso, J. Am. Chem. Soc. 2022, 144, 2765. 162 N. V. Tkachenko, S. Scheiner, ACS Omega 2019, 4, 6044. 163 J. Areephong, N. Ruangsupapichart, T. Thongpanchang, Tetrahedron Lett. 2004, 45, 3067. 164 C. A. Guido, F. Zinna, G. Pescitelli, J. Mater. Chem. C, 2023, 11, 10474. 165 L. Zhang, I. Song, J. Ahn, M. Han, M. Linares, M. Surin, H.-J. Zhang, J. H. Oh, J. Lin, Nat Commun. 2021, 12, 142. 166 M. Cei, L. Di Bari, F. Zinna, Chirality 2023, 35, 192. Chapter 4 129 Thereafter, the camphorsulphonyl group is removed by basic hydrolysis, affording enantiopure BINOL-based polyarenes M-25 and P-25. Finally, the Scholl reaction providing the helical bilayer nanographene oxa[9]HBNG occurs enantiospecifically. The configurationally stable atropisomers M-25 and P-25 are transformed into helical fully conjugated nanographenes due to chirality transference, to obtain enantiopure M-oxa[9]HBNG and P-oxa[9]HBNG, respectively. Scheme 4.2. Enantiospecific synthesis of M-oxa[9]HBNG and P-oxa[9]HBNG from M-25 and P-25, respectively, after chemical resolution of the enantiomers. Chapter 4 130 The cyclodehydration of BINOL derivatives for the preparation of oxahelicenes is a commonly used approach.167 However, the reported reaction conditions in presence of Brønsted acids (p-toluene sulfonic acid and trifluoromethane sulfonic acid) require high temperatures, commonly reflux of toluene, chlorobenzene or xylene, and long reaction times, resulting in lower enantiomeric excesses (e.e.) due to racemization.168 Alternatively, Sasai et al. reported a one-pot enantioselective oxidative coupling followed by intramolecular cyclization using a complex chiral vanadium catalyst, affording oxa[9]helicene with a maximum 78% ee.169 Therefore, the mild reaction conditions (−30 ºC for 40 min) used to perform the simultaneous cyclodehydrogenation and cyclodehydration, in presence of DDQ and triflic acid, is a suitable and affordable approach towards the enantiospecific synthesis of a new oxa-helical bilayer nanographene. The mechanism for the cyclodehydration of BINOL-like structures to afford furane-embedded helicenes under acidic conditions is known to occur through the formation of cation intermediates,163 as shown in Scheme 4.3. The tautomerization of one naphthol group in M-25′ under acidic conditions provides the electrophilic oxonium intermediate INT1. The intramolecular nucleophilic addition of the proximal hydroxy group to the Si face of the oxonium provides the five-membered cyclic hemiacetal INT2. Eventually, upon dehydration, the system is rearomatized and enantiopure M-oxahelicene is obtained. The face for the attack in the enantiospecific cyclization from INT1 to INT2 is determined by the fixed spatial orientation of the naphthol groups, which depends on the enantiomer of polyarene 25. Scheme 4.3. Cyclodehydration mechanism through cation intermediates. The enantiomeric excess of the M and P-oxa[9]HBNG were evaluated by chiral HPLC (column Chiralpack IC 5μm, 98% hexane 2% toluene:isopropanol 4:1, Appendix 4), showing an e.e.> 99% for M-oxa[9]HBNG and e.e.= 99% for P- oxa[9]HBNG. Considering the enantiospecific mechanism of the reaction, the enantiomeric excess depends on the enantiopurity of the starting materials. The hydrolisis conditions to afford P and M-25 are mild enough to prevent racemization, therefore the proper separation of the diastereomers by silica gel column determines 167 K. Nakanishi, D. Fukatsu, K. Takaishi, T. Tsuji, K. Uenaka, K. Kuramochi, T. Kawabata, K. Tsubaki, J. Am. Chem. Soc. 2014, 136, 7101. 168 S, Hossain, M. Akter, M. Karikomi, Tetrahedron Lett. 2021, 69, 152948. 169 M. Sako, Y. Takeuchi, T. Tsujihara, J. Kodera, T. Kawano, S. Takizawa, H. Sasai, J. Am. Chem. Soc. 2016, 138, 11481. Chapter 4 131 the e.e. of the final bilayer nanographene. The purity of the diastereomers was determined by 500MHz 1H-NMR, revealing an excellent isolation (purity > 95%). The final products, as well as the intermediates, were completely characterized by NMR, mass spectrometry and FT-IR. Polyarene derivatives R,M-26a and R,P-26b show clear differences by NMR, as expected from diastereomers. The substitution with only one camphorsulfonyl group entails a great asymmetry in the spectra. As represented in Figure 4.1, both diastereomers present ten singlets, between 1.11 and 1.00 ppm, corresponding to the tert-butyl groups, which shows the symmetry loss after the functionalization, polyarene 25 exhibits signals corresponding to one half of the molecule (Appendix 5. Spectra). Among the large number of signals corresponding to pure diastereomers R,M-26a and R,P-26b, three can be highlighted to emphasize the shifting differences, being also useful to evaluate the purity of both diastereomers (see also Appendix 5. Spectra for further comparison). On the one hand, the aliphatic singlets, at 0.84 and 0.58 ppm for R,M-26a, and at 0.81 and 0.60 ppm for (R,P)-26b, which correspond to the diasterotopic methyl groups in the bridge positions of the camphorsulphonyl substituents. On the other hand, the aromatic singlet at 7.02 ppm (R,M-26a) and 7.07 ppm (R,P-26b), corresponding to the 8 or 8′ position in the BINOL core. Figure 4.1. 1H-NMR spectra corresponding to diastereomers (R,M)-26a (top) and (R,P)-26b (bottom). The 1H-NMR characterization of oxa[9]HBNG exhibits eleven singlets and two doublets at low field corresponding to the aromatic protons, and five singlets at high field corresponding to the tert-butyl groups, shown in Figure 4.2. The number of 1H-NMR 26a 1H-NMR 26b 0 .8 4 0 .5 8 0 .8 1 0 .6 0 7 .0 2 7 .0 7 Chapter 4 132 signals reveals the symmetry of the structure due to a C2 axis. As can be seen, the expected phenol signal between 3.50 and 4.00 ppm does not appear in the spectrum, which is evidence of closure to form the furan-like cycle. In addition, mass spectrometry experiments show the loss of 42 units from 25 (exact mass, 1759.1040) to oxa[9]HBNG (exact mass, 1716.8948), which agrees to the formation of 12 C−C bonds (24 units), and the loss of 18 units due to cyclodehydration to form a new C−O bond. Figure 4.2. 1H-NMR spectrum corresponding to oxa[9]HBNG. Furthermore, the structure of diastereomer R,P-26b, as well as both enantiomers of nanographene oxa[9]HBNG, were unequivocally characterized by single-crystal X- ray diffraction. As detailed in Chapter 2.1, atropisomers are assigned looking along the chirality axis that contains the single bond with hindered rotation. Thus, placing the naphthalene unit substituted with the camphorsulphonyl group at the front (Cahn- Ingold-Prelog rules), the BINOL derivatives are P if the order follows clockwise rotation, and M if the rotation is anticlockwise. Thus, diastereomer 26b presents the P-BINOL core, as represented in Figure 4.3. Chapter 4 133 Figure 4.3. Single-crystal X-ray structure of 26b, P-BINOL. The analysis of nanographene oxa[9]HBNG by single-crystal X-ray diffraction showed a partially overlapped bilayer. The incorporation of the oxa-five-membered ring to the [9]helicene entails the opening of the inner rim of the helicene, decreasing the number of turns. Thus, 22 benzene rings are involved in the intramolecular π-π interactions vs. 26 rings in the case of the totally overlapped structure [9]HBNG, (Figure 3.9, Chapter 3.3.1). Enantiomerically pure samples of nanographene oxa[9]HBNG were crystallized, thus confirming the expected configuration of the helicenes considering the enantiospecific mechanism for the cyclization (Scheme 4.3). Thus, polyarene M-25 provides M-oxa[9]HBNG and polyarene P-25 provides P- oxa[9]HBNG (Figure 4.4). Figure 4.4. Crystal structures of M-oxa[9]HBNG and P-oxa[9]HBNG from their respective enantiopure samples. 4.3.2 Electrochemical properties The electrochemical properties of oxa[9]HBNG were evaluated by cyclic voltammetry in a 0.1 M solution of tetrabutylammonium hexafluorophosphate in toluene/acetonitrile 4:1 using glassy carbon as working electrode, a reference electrode of Ag/AgNO3, and a platinum wire as counter electrode. The oxidation and reduction potentials vs. Fc/Fc+ are summarized in Table 4.1, along with the potentials of [9], [10] and [11]HBNG for comparison purpose. As shown in Figure 4.5, * P * P*M Chapter 4 134 oxa[9]HBNG presents two quasi-reversible oxidation waves and two quasi-reversible reduction waves. The potential of the first oxidation wave E1 1/2ox= 0.37 V is slightly higher than the first oxidation potential of [9]HBNG E1 1/2ox= 0.35 V. Comparing the oxidation waves with those of nanographene 3.8 E1 1/2ox= 0.9 V (in CH2Cl2), oxa[7]helical nanographene reported by Jux et al. in which the HBC layers do not show intramolecular π-π interactions,137 oxa[9]HBNG shows a significantly stronger donor character. As described for [9], [10] and [11]HBNG, the variation of the π- extension of the helicene (oxa[7]helicene Jux et al. vs oxa[9]HBNG) does not necessarily modulates the donor character in layer-overlapping structures. Therefore, it can be concluded that the contribution of heteroatom in the helicene has a lower impact over the oxidation potential. Therefore, the strong donor character of oxa[9]HBNG is attributed to the overlapping degree between the layers. Oxa[9]HBNG exhibits less layer overlapping (22 rings involved in the π-π interactions) than [9]HBNG (26 rings involved in the π-π interactions), which decreases the mixed valence band effect of the bilayer to stabilize the cation and radical cation species between the layers, eventually increasing the oxidation potential to 0.37 V. Figure 4.5. Cyclic voltammogram of oxa[9]HBNG showing the half-wave redox potentials vs. Fc/Fc+ in a 1 M solution of tetrabutylammonium hexafluorophosphate in toluene/acetonitrile 4:1. However, the oxygen embedded in the helicene seems to be more important regarding the electron acceptor character of oxa[9]HBNG. As described in Chapter 3.3.2, the LUMO orbitals of [9], [10] and [11]HBNG, are mainly located in the helicene backbones, Figure 3.11. In the case of oxa[9]HBNG, the electron-donor character of both the helicene (due to the electron-donor character of the oxygen), and the HBC layers, entails the destabilization of the LUMO orbital. Thus, the potential of the first reduction wave increases (in absolute value) to E1 1/2red= −2.27 V, revealing a weaker electron acceptor character vs. [9], [10] and [11]HBNG (Table 4.1), and even tBu- HBC (E1 1/2red= −2.24 V). 1.0 0.5 0.0 -0.5 -1.0 -1.5 -2.0 -2.5 x x x x -2.52 -2.27 0.66 C u rr e n t Potential [V] vs, Fc/Fc + CV oxa[9]HBNG 0.37 Chapter 4 135 Table 4.1. Half-wave redox potentials of oxa[9]HBNG, [9]HBNG, [10]HBNG and [11]HBNG. Compound E1 1/2ox (V) E2 1/2ox (V) E1 1/2red (V) E2 1/2red (V) oxa[9]HBNG 0.37 0.66 –2.27 –2.52 [9]HBNG 0.35 0.59 –2.18 –2.46 [10]HBNG 0.46 0.67 –2.23 –2.55 [11]HBNG 0.52 0.69 –2.22 –2.46 4.3.3 Photophysical Properties The photophysical properties of oxa[9]HBNG were evaluated by UV-vis absorption and emission measurements in chloroform (summarized in Table 4.2). Resembling bilayer nanographenes [9], [10] and [11]HBNG, oxa[9]HBNG displays non- structured broad bands in the ultraviolet region of the absorption spectrum (344, 371, 398 nm very similar to those of the previously described HBNGs, and more structure weak bands in the visible region (450, 476, 498 nm) (Figure 4.6, yellow line). Figure 4.6. Absorption (solid lines) and emission (shaded curves) spectra of oxa[9]HBNG, [9]HBNG, and [11]HBNG in chloroform. The emission spectrum of oxa[9]HBNG shows well-structured bands with maxima at 500, 537 and 580 nm, which contrasts with the broad bands display by [9], [10] and [11]HBNG. As illustrated in Figure 4.6, the emission bands (shaded curves) of oxa[9]HBNG are blue-shifted with respect to the previously described carbon-based HBNGs, and are more comparable to those of tBu-HBC. These observations agree with the cyclic voltammetry data previously described. The fluorescence quantum yield of oxa[9]HBNG resulted of Φem= 0.26, determined by comparison with the emission spectra of fluoresceine in ethanol (Φem= 0.97) at 432 nm. The incorporation of the oxygen atom in the helicene, modifies the structure and entails the modification of the LUMO orbital, which is not farther localized in the helicene moiety, as for oxa[9]HBNG [9]HBNG [11]HBNG t Bu-HBC 300 350 400 450 500 550 600 650 700 N o rm a liz e d o p ti c a l d e n s it y Wavelength (nm) Chapter 4 136 carbon-based HBNGs. This modification involves the variation of the photophysical properties, the emission bands are no longer excimer-like, and the quantum yield follows the expected order, in which a greater band gap (E0-0, Table 4.2) entails greater quantum yield. Table 4.2. Data of absorption and fluorescence spectra of HBNGs. Compound λabs max /nm (ε/ M−1·cm−1) λem max /nm Φem E0-0/eV oxa[9]HBNG 344 (66,647); 371 (141,645); 398 (71,296); 449 (23,690); 476 (18,545); 498 (8,102) 500, 537, 580 0.30 2.55 [9]HBNG 349 (94,600); 373 (136,000); 392 (86,300); 435 (20,600); 502 (6,200); 542 (4,700) 575, 611 0.22 2.35 [10]HBNG 360, 377, 415, 446, 462, 494 543, 567 0.10 2.49 [11]HBNG 339, 377, 415, 444, 491 528, 551 0.11 2.58 4.3.4 Chiroptical properties The electronic circular dichroism (CD) has been recorded in chloroform for both enantiomerically pure nanographenes M-oxa[9]HBNG and P-oxa[9]HBNG. As illustrated in Figure 4.7, the spectra of the enantiomers (P-oxa[9]HBNG, brown solid line, and M-oxa[9]HBNG, orange solid line) show specular image signals and two main regions with prominent Cotton effects. One intense broad band from the ultraviolet region to the visible region (300 nm to 470 nm), and one of lower intensity at lower energy in the visible region (480 nm). The absorption dissymmetry factors represented with dashed lines in Figure 4.7, are gabs= +3.7·10−3 at 487 nm for M- oxa[9]HBNG, and gabs= −1.8·10−3 at 487 nm for P-oxa[9]HBNG. Figure 4.7. CD spectra and gabs values of M-oxa[9]HBNG and P-oxa[9]HBNG. 300 325 350 375 400 425 450 475 500 -30 -20 -10 0 10 20 30 CD, M-oxa[9]HBNG CD, P-oxa[9]HBNG Wavelength (nm) -0.05 -0.04 -0.03 -0.02 -0.01 0.00 0.01 0.02 0.03 0.04 0.05 C D ( m d e g ) g a b s gabs, M-oxa[9]HBNG gabs, P-oxa[9]HBNG Chapter 4 137 4.4 Conclusions To conclude, in this chapter a new approach for the enantiospecific synthesis of helical bilayer nanographenes by an easily scalable methodology, is described. The synthesis starts from commercially available 7-bromo-2-naphtol that, in presence of copper (II), yields axially chiral dibromoBINOL. Taking advantage of the stable axial chirality and reactivity of the obtained BINOL-based structures, a chemical resolution was carried out by functionalizing the BINOL-based polyarene with enantiopure camphorsulphonyl chloride. After a hydrolysis step under mild conditions, enantiopure BINOL-based polyarenes P- and M-25 were obtained, which under Scholl reaction conditions lead to the corresponding enantiopure P- and M- oxa[9]HBNG. During the Scholl oxidation two different types of reactions take place, the expected cyclodehydrogenation (providing the graphene like structure), and the enantiospecific cyclodehydration of the phenol groups yielding the furane embedded helicene (chirality transference from axial to helical). The separation of the diastereomers by silica gel column afforded the diastereomers with an isomeric excess > 95% (by NMR), the performance of this separation determines the enantiomeric excess of the final nanographene (e.e.≥ 99% by chiral HPLC). All structures were characterized by the common techniques, and single-crystal X-ray diffraction confirmed the configuration of the BINOL core in diastereomeric polyarene R,P-26b (P-BINOL) and the corresponding enantiomer obtained after hydrolisis and Scholl reaction, P-oxa[9]HBNG. Furthermore, the crystal structure of nanographene oxa[9]HBNG showed the great overlapping between the layers, being 22 rings involved in the intramolecular π-π interactions. The great overlapping of the structure explains the variation of the donor character. The influence of the electron-donor character of oxygen in the helicene backbone has a lower impact on the redox properties than the influence of the overlapping degree between the layers. Comparing oxa[9]HBNG with oxa[7]helical nanographene, reported by Jux et al., in which the HBC layers do not show intramolecular π-π interactions (E1 1/2ox= 0.9 V in DCM),137 and with the corresponding all-carbon nanographene [9]HBNG (E1 1/2ox= 0.35 V), the observed first oxidation potential of oxa[9]HBNG (E1 1/2ox= 0.37 V) is mainly influenced by the overlapping degree. The slightly higher potential than that of the corresponding carbon-based structure agrees with the lower overlapping degree (22 vs 26 rings involved in the π-π interactions in oxa[9]HBNG and [9]HBNG respectively). The influence of the electron donor character of the oxygen-containing helicene is more reflected in the acceptor character. In all carbon structures [9], [10] and [11]HBNG the LUMO is located in the helicene feature. However, the electron donor character of the oxa[9]helicene leads to a higher energy LUMO orbital. Thus, the first reduction potential for oxa[9]HBNG (E1 1/2red= −2.27 V) is closer to that of tBu-HBC (E1 1/2red= −2.24 V). In addition, the emission spectra show well-structured bands suggesting the lower excimer-like contribution and the delocalization of the LUMO in the layers of the structure instead Chapter 4 138 of delocalized in the helicene moiety. Finally, the chiroptical properties of M- and P- oxa[9]HBNG have been studied by means of electronic circular dichroism, from which gabs values of +3.7·10−3 at 487 nm for M-oxa[9]HBNG, and −1.8·10−3 at 487 nm for P-oxa[9]HBNG, have been calculated. Experimental section Experimental section 141 EXPERIMENTAL SECTION Unless otherwise noted, all materials including solvents were obtained from commercial suppliers and used without further purification. Unless otherwise noted, all reactions were performed with dry solvents (dried by filtration through alumina according to the method described) and under an atmosphere of argon in dried glassware with standard vacuum-line techniques. All work-up and purification procedures were carried out with reagent-grade solvents in air. Silica column chromatography was conducted with Scharlau 40-60 μm silica gel. Analytical thin-layer chromatography (TLC) was performed using Silica gel 60 F254-coated aluminum sheets (Merck). Flash chromatography was performed on Silica gel 60 (0.040-0.063 mm, Merck). The developed chromatogram was analyzed by solution of Ce(SO4)2·4 H2O (1%) and H3P(Mo3O10)4 (2%) in sulfuric acid (10%) and UV lamp (254 nm and 365 nm). Microwave reactions were performed in an Anton-Parr Monowave 300 microwave or in an Anton-Parr Monowave 400 microwave reactor. 1H NMR spectra were recorded at 700 (Bruker AVIII), 500 (Bruker AV), 400 (Bruker AVIII) or 300 (Bruker AVIII) MHz and 13C NMR spectra were recorded at 176, 126 or 101 (Bruker AV) MHz. Chemical shifts for 1H NMR and 13C NMR are expressed in parts per million (ppm). For referencing of NMR spectra, the residual solvent signal (CDCl3 δ 7.26 ppm for 1H and δ 77.16 ppm for 13C, C6D6 δ 7.16 ppm for 1H and δ 128.06 ppm for 13C, Tetrachloroethane-d2 δ 6.00 ppm for 1H and δ 73.78 ppm for 13C, or Methylene Chloride-d2 δ 5.32 ppm for 1H and δ 54.00 ppm for 13C) was used. Data are reported as follows: chemical shift, multiplicity (s= singlet, d= doublet, dd= doublet of doublets, t= triplet, q= quartet, m= multiplet, bs= broad singlet), coupling constant (Hz), and integration. MALDI-ToF matrices were trans-2-[3-(4-tert-butylphenyl)-2-methyl-2-propenylid ene]-malononitrile (DCTB) or 1,8-dihidroxi-9,10-dihydroanthracen-9-ona (dithranol) and mass analysis were performed in a Bruker Ultraflex II using an LTB MNL 106 laser source. The APCI mass spectra were recorded using an LTQ Orbitrap XL (Thermo Fisher Scientific) hybrid mass spectrometer equipped with an APCI ion source. The APCI vaporizer and heated capillary temperatures were set to 400 °C and 200 °C, respectively; the corona discharge current was 3.5 μA. Nitrogen served both as the sheath and auxiliary gas at flow rate 55 and 5 arbitrary units, respectively. The ionization conditions were the same for low‐resolution as well as high‐resolution experiment. The HR spectra were acquired at a resolution of 100 000. IR spectra were recorded on a FT-IR Nicolet Magna 750 spectrometer and in KBr cell in CHCl3 solution on Nicolet 6700 FT-IR spectrometer (Thermo Fisher Scientific, USA) equipped with a standard mid-IR source, a KBr beam-splitter, and a DTGS detector and with a cell compartment purged by dry nitrogen. Experimental section 142 Electrochemical measurements were performed using a standard one-compartment, three-electrode electrochemical cell connected to an electrochemical analyzer (Metrohm Autolab). The working electrode was a glassy carbon electrode (3 mm diameter) that was freshly polished with a suspension of Al2O3 in distilled water and sonically rinsed with acetone before each measurement. Silver (Ag/0.1 M AgNO3 in CH3CN) and platinum wires were used as reference and counter electrodes, respectively. Electrochemical grade (Aldrich) tetrabutylammonium hexafluro- phosphate 0.1 M in toluene:acetonitrile was used as supporting electrolyte. All measurements were conducted under dry argon. Solutions were saturated with argon for deoxygenation and to maintain an argon blanket for at least 10 minutes prior to each measurement. All measurements are referenced to Fc/Fc+ added as internal reference. UV-vis absorption spectra of nanographenes 10, 13, 17 and oxa[9]HBNG were recorded using a Cary 50 Conc spectrophotometer and quartz cells of 1.0 cm optical path length. Emission spectra of nanographenes 10, 13 and 17 were acquired in a Fluorolog TC-SPC HORIBA-JOVIN YVON spectrofluorometer in quartz cells of 1.0 cm optical path length. Samples were dissolved in chloroform (Carlo Erba, for analysis, ≥99%) and the solutions were in the 1–5 μM range. UV-vis absorption spectra of [9], [10] and [11]HBNG were recorded on a UV-vis spectrophotometer (Varian Cary 50) using 1 cm quartz cuvette (Suprasil). Room-temperature fluorescence spectra were acquired on a spectrofluorimeter (Horiba FluoroLog 3) equipped with a high-pressure Xenon lamp and a Hamamatsu R928P photomultiplier tube. Fluorescence spectra of [9], [10], [11]HBNG and oxa[9]HBNG were corrected for the characteristics of the lamp source and of the detection system. Fluorescence lifetimes (τ) were measuring via time-correlated single photon counting (TCSPC) measurements employing an Acton SP2500 spectrometer equipped with a PMA 06 photomultiplier (PicoQuant) and a HydraHarp-400 TCSPC event timer with 1 ps time resolution. The excitation source was a 405 nm picosecond pulsed diode laser (LDH- D-C-405, PicoQuant) driven by a PDL828 driver (PicoQuant) with FWHM < 70 ps. Fluorescence decays were analyzed using PicoQuant Fluofit v4.6.5 data analysis software. Spectroelectrochemical Measurements. In situ UV-vis-NIR spectroelectrochemical oxidations of [9], [10] and [11]HBNG were conducted on a Varian Cary 5000 UV- vis-NIR spectrophotometer. A C3 epsilon potentiostat from BASi was used for the electrolysis using a thin layer cell from a demountable omni cell from Specac. In this cell a three electrodes system was coupled to conduct in situ spectroelectrochemistry. A Pt gauze was used as the working electrode, a Pt wire was used as the counter electrode, and an Ag wire was used as the pseudo-reference electrode. The spectra were collected at constant electrolysis potential and the potentials were changed in interval of 10 mV. The electrochemical medium used was 0.1 M tetrabutylammonium Experimental section 143 hexafluorophosphate, Bu4NPF6, in fresh distilled dichloromethane, at room temperature with sample concentrations of 10−3 M. Vibrational Raman Spectroscopy. The 1064 nm Raman spectra of tBu-HBC, [10]HBNG and [11]HBNG in solid state at room temperature were measured using the RAMII FT-Raman module of a VERTEX 70 FT-IR spectrometer (Bruker). A continuous–wave Nd–YAG laser working at 1064 nm was employed for excitation, at a laser power in the sample not exceeding 10 mW. Raman scattering radiation was collected in a back–scattering configuration with a standard spectral resolution of 4 cm−1. 10000 scans were averaged for each spectrum. The 325 nm Raman spectrum of [9]HBNG in solid state at room temperature was obtained with a JASCO NRS-5000 dispersive laser Raman spectrometer with a confocal microscope and equipped with a Peltier cooled CCD detector operating at –60 ºC. Raman spectrum was recorded by averaging spectra during 16 min with a resolution of 1-2 cm–1. HPLC analysis of nanographene 17a was collected on an Agilent Technologies 1260 Infinity HPLC system using Regis Pirkle-type covalently modified silica column (4.6 × 250 mm) Whelk-O2 (R,R). Injection volume was 0.7 μL, flow rate was 1 mL/min and eluent was 98:1:1 hexane:isopropanol:tetrahydrofuran. Peaks were confirmed by comparison of their UV-Vis spectra to originate from the enantiomers. HPLC analyses of [9], [10] and [11]HBNG were performed on Knauer Smartline 1000/2500 isocratic system with polarimetric detector (Chiralyser MP, IBZ Messtechnik) connected to Clarity chromatography station. Semipreparative HPLC resolutions were performed on PuriFlash 5.250 (Interchim) quaternary gradient HPFC/HPLC system equipped with DAD detector. Optical rotations were measured using an Autopol IV instrument (Rudolph Research Analytical). HPLC analyses of oxa[9]HBNG were performed on JASCO LC-4000 series, dotted with a JASCO PU-4180 RHPLC pump, a JASCO AS- 4050 autosampler, and a JASCO CO-4061 column oven; all connected to a JASCO CD-4095 UV-vis and CD detector. The HPLC analytical column used was Chiralpack IC (5 μm, 25 cm x 4.6 mm ID). The ECD spectra of [9], [10] and [11]HBNG were measured on a Jasco 1500 spectropolarimeter (JASCO International Co. Ltd.) equipped with a fluorescence emission monochromator (FMO522) and separate fluorescence emission detector (FDT-538). The ECD spectra were measured over a spectral range of 210 nm to 550 nm in tetrahydrofuran (10-4 M solutions). Measurements were made in a quartz cell with a 0.1 cm path length using a scanning speed of 20 nm/min, a response time of 4 seconds and standard instrument sensitivity. After a baseline correction, the CD spectra were expressed in terms of differential molar extinction (Δε). The ECD spectra of oxa[9]HBNG were measured on a JASCO J-1500 CD Spectrometer, over a spectral range of 250 nm to 550 nm in chloroform (ca. 10−5 M solutions). Measurements were made in a quartz cell with a 1 cm path length using a scanning speed of 50 nm/min, a response time of 4 seconds and standard instrument sensitivity. Experimental section 144 The circularly polarized luminescence (CPL) measurements of [9], [10] and [11]HBNG were performed using a home-built CPL spectrofluoropolarimeter (constructed with the help of the JASCO Company). The samples were excited using a 90° geometry with a Xenon ozone-free lamp 150 W LS. The following parameters were used: emission slit width ≈ 10 mm, integration time= 8 sec, scan speed= 50 nm/min, accumulations= 5. The concentration of all the samples was ca. 10−6 M. Quantum Chemistry Calculations. Density Functional Theory (DFT) quantum chemical calculations were performed with the Gaussian’16 suite of programs. The Becke’s three parameter (B3) gradient-corrected exchange functional was used, and the non-local correlation was provided by the Lee-Yang-Parr (LYP) expressions. -To achieve an improved description of long-range interactions we used the Coulomb- attenuating hybrid method (CAM-B3LYP), which includes the Hartree-Fock and the Becke exchanges as a variable ratio depending on the intermolecular distance. Structural optimizations and spectroscopic features were calculated using the split- valence 3-21g(d) basis set. Raman spectra were obtained from the calculated DFT intensities, and the vibrational wavenumbers uniformly scaled by 0.96. Every band was represented by a Gaussian function of 10 cm−1 half-height width. Electronic excitation energies were obtained by using the time dependent DFT (TDDFT) formalism for which up to the fifty low-lying energy states were considered. Experimental section 145 Chapter 1 9,10-Bis[(4-(tert-butyl)phenyl)ethynyl] anthracene (4). To a dry 10 mL microwave reactor vial under argon atmosphere and provided with a magnetic stir bar, CuI (0.180 mmol, 34 mg, 0.4 equiv.), Pd(dppf)Cl2 (0.089 mmol, 73 mg, 0.2 equiv.), 9,10- dibromoanthracene 2 (0.446 mmol, 150 mg, 1 equiv.), 4-(tert-butyl)phenylacetylene 3 (1.120 mmol, 180 mg, 2.5 equiv.), 4.6 mL of THF anhydrous (previously deoxygenated by freeze-pump-thaw) and 1.2 mL of freshly distilled DIPA were added. The reaction was carried out for 3 hours at 120 ºC and 600 rpm in a microwave reactor. The crude was extracted in DCM washing with NH4Cl (20 mL), water (20 mL) and brine (20 mL). The resultant crude was filtered through a silica gel plug with dichloromethane and solvent was removed. The product was dissolved in the minimum amount of CHCl3 and then methanol was added to obtain a precipitate that was filtered to reach to a yellow solid (185 mg, 85%). 1H (300 MHz, CDCl3) δ= 8.72 − 8.67 (m, 4H), 7.72 (d, J= 8.5 Hz, 4H), 7,66 − 7.60 (m, 4H), 7.48 (d, J= 8.5 Hz, 4H), 1.38 (s, 18H). FT-IR (cm−1) 3084, 3059, 3032, 2954, 2949, 2901, 2870, 2858, 2319, 1910, 1742, 1664, 1617, 1513, 1502, 1456, 1434, 1398, 1361, 1260, 1105, 1024, 829, 765, 639, 559. HRMS found: 490.2671. Calculated for C38H34, 490.2661.170 9,10-Bis[penta(4-tert-butyl(phenyl))phenyl] anthracene 6a. To a 30 mL dry microwave reactor vial, provided with a magnetic stir bar were added, 2,3,4,5-tetrakis[4-(tert-butyl)phenyl]cyclopenta-2,4 -dien-1-one 5 (0.569 mmol, 346 mg, 3.1 equiv.) and dialkyne 4 (0.183 mmol, 90 mg, 1 equiv.). The reaction was carried out in a microwave reactor at 280 ºC and 600 rpm for 40 minutes, repeating these conditions for a total of three sequential times. The crude was purified through aluminum oxide chromatography column using as eluent a gradient from hexane to hexane:DCM 90:10 affording the product 6a as a yellow solid (170 mg, 57 %). 1H NMR (300 MHz, CDCl3) δ= 7.63 − 7.59 (m, 4H), 6.83 (d, J= 8.5 Hz, 4H), 6.77 − 6.73 (m, 12H), 6.70 (d, J= 8.5 Hz, 4H), 6.61 (d, J= 8.4 Hz, 8H), 6.49 (d, J= 8.6 Hz, 8H), 6.30 (d, J= 8.4 Hz, 8H), 1.12 (s, 18H), 1.08 (s, 36H), 0.99 (s, J= 3.3 Hz, 36H). 13C NMR (75 MHz, CDCl3) δ= 147.4, 147.2, 146.9, 143.0, 141.3, 140.9, 138.6, 138.5, 137.8, 135.7, 134.5, 131.5, 131.3, 130.9, 129.6, 127.8, 123.3, 123.1, 123.0, 122.3, 34.2, 34.1, 34.1, 31.2. FT-IR (cm-1) 3080, 3051, 3027, 2958, 2905, 2865, 1779, 1662, 1602, 1511, 1462, 1390, 1361, 1268, 1116, 1020, 831, 567. HRMS found: 1651.0783. Calculated for C126H138, 1651.0799. 170 W. Fudickar, T. Linker, J. Am. Chem. Soc. 2012, 134, 15071. Experimental section 146 Syn/anti spirocompound 7. In a 25 mL dry Schlenk, under argon atmosphere and provided with a magnetic stir bar, 6a (0.012 mmol, 20 mg, 1 equiv.) and AgPF6 (0.606 mmol, 153 mg, 50 equiv.) were dissolved in 5 mL of anhydrous DCE. The solution was heated to 85 ºC and a solution of FeCl3 (0.606 mmol, 77 mg, 50 equiv.) in 2 mL of nitromethane was added dropwise under argon bubbling. The reaction was carried out at 85 ºC under Argon bubbling and stirring for 4 hours. After that time methanol was added to the reaction Schlenk and solvent was removed under reduced pressure. The crude was purified through neutral aluminum oxide preparative plate with CS2, affording 7 as a white solid (12 mg, 61%). 1H NMR (700 MHz, Methylene Chloride-d2) δ= 7.42 – 7.27 (m), 6.96 – 6.88 (m), 6.81 (dd, J= 8.5, 1.8 Hz), 6.74 – 6.65 (m), 6.53 – 6.47 (m), 6.29 – 6.25 (m), 6.11 (d, J= 8.4 Hz), 6.07 – 6.02 (m), 1.36 (s), 1.35 (s), 1.34 (s), 1.13 (s), 1.01 (s), 1.03 (s), 1.03 (s), 0.91 (s), 0.91 (s), 0.90 (s). 13C NMR (176 MHz, Methylene Chloride-d2) δ= 159.8, 151.6, 150.3, 149.8, 148.4, 147.7, 146.5, 142.5, 142.1, 140.3, 139.6, 139.3, 138.3, 138.0, 137.9, 135.5, 135.4, 133.1 (bs), 132.2 (bs), 131.7 (bs), 130.7 (bs), 124.2, 123.7, 123.5, 123.2, 122.9, 63.8, 35.0, 35.0, 34.5, 34.4, 34.2, 31.7, 31.5, 31.4, 31.4. FT-IR (cm-1) 3072, 3027, 2962, 2929, 2901, 2860, 1735, 1604, 1517, 1475, 1362, 1270, 1115, 1022, 832. HRMS found: 1649.0642. Calculated for C126H136, 1649.0642. Spirocompound 8. In a 10 mL dry Schlenk, under argon atmosphere and provided with a magnetic stir bar, 6a (0.012 mmol, 20 mg, 1 equiv.) was dissolved in 2 mL of anhydrous DCM. The solution was cooled to –30 ºC and a solution of FeCl3 (0.606 mmol, 77 mg, 50 equiv.) in 0.6 mL of nitromethane was added dropwise under argon bubbling. The reaction was carried out at –30 ºC under Argon bubbling and stirring for 1 hour. After that time methanol was added to the reaction Schlenk and solvent was removed under reduced pressure. The crude was heated in ethanol, after cooling a yellow solid was filtered, providing 8 as a yellow solid (8 mg, 40%) HRMS found: 1643.0344. Calculated for C126H130, 1643.0173. Spironanographene syn 10. To a 100 mL dry round-bottom flask, under argon atmosphere and provided with a magnetic stir bar, 1a (0.012 mmol, 20 mg, 1 equiv.) and DDQ (0.363 mmol, 82 mg, 30 equiv.) were added and dissolved in 30 mL of anhydrous DCM. The solution warmed to 40 ºC with stirring. Then, TfOH (4.85 mmol, 0.730 g, 400 equiv.) was added dropwise under argon bubbling. The reaction was carried out at 40 ºC under Argon bubbling and stirring for 2 hours. After that time a HNaCO3 saturated solution was added to the reaction flask and the crude was extracted with DCM and washed twice with a Experimental section 147 HNaCO3 saturated solution (50 mL) and twice with brine (50 mL). From this step the organic phase was dried with MgSO4 and filtered, the solvent was removed under reduced pressure. Then, the crude was filtered through a small silica gel plug with DCM. The solvent was removed under reduced pressure and the product was purified by a semipreparative silica gel plate in hexane:DCM 90:20, affording syn 10 as a yellow solid (4 mg, 21%). Syn 10 1H NMR (500 MHz, CDCl3) δ= 9.34 (s, 1H), 9.28 (s, 1H), 9.26 (s, 1H), 9.23 (s, 1H), 9.07 (s, 1H), 8.82 (s, 1H), 8.65 (s, 1H), 8.23 (d, J= 2.0 Hz, 1H), 7.99 (d, J= 8.6 Hz, 1H), 7.78 (s, 1H), 7.03 – 6.99 (m, 1H), 6.68 – 6.65 (m, 1H), 6.64 – 6.57 (m, 2H), 5.80 (dd, J = 8.6, 2.0 Hz, 1H), 1.86 (s, 9H), 1.85 (s, 9H), 1.77 (s, 9H), 1.50 (s, 9H), 0.72 (s, 9H). 13C NMR (126 MHz, CDCl3) δ= 158.6, 152.1, 149.7, 149.5, 148.6, 148.2, 146.1, 139.3, 137.6, 136.4, 133.2, 131.5, 131.1, 130.9, 130.8, 130.4, 130.3, 130.2, 130.2, 130.1, 128.5, 127.9, 127.7, 127.6, 127.6, 124.8, 124.4, 124.3, 123.9, 123.0, 122.9, 122.1, 119.3, 119.2, 118.8, 118.8, 118.7, 118.6, 118.5, 118.1, 117.9, 69.9, 36.3, 36.0, 36.0, 35.7, 34.0, 32.6, 32.3, 32.3, 32.1, 30.9. FT- IR (cm-1) 3067, 2957, 2921, 2865, 1751, 1689, 1601, 1476, 1363, 1251, 1202, 1096, 1023, 870, 620. HRMS found 1632.9375. Calculated for C126H120 1632.9390. 9,10-Bis[(4-(tert-butyl)phenyl)ethynyl]octafluoro anthracene (12). To a dry 30 mL microwave reactor vial under argon atmosphere and provided with a magnetic stir bar, CuI (0.020 mmol, 4 mg, 0.4 equiv.), Pd(PPh3)2Cl2 (0.010 mmol, 7 mg, 0.2 equiv.), 9,10-dichloro-1,2,3,4,5,6,7,8-octafluoroanthracene 11 (0.051 mmol, 20 mg, 1 equiv.), 4-(tert-butyl)phenylacetylene 3 (0.153 mmol, 24 mg, 3 equiv.), 4 mL of deoxygenated (argon bubbling for 30 minutes) freshly distilled triethylamine. The reaction was carried out for 3h and 30 minutes at 80 ºC and 300 rpm in a microwave reactor. Solvent was removed from the crude under reduced pressure then dissolved in DCM and filtered through a silica gel plug column using dichloromethane as eluent. The solvent was removed under reduced pressure. The obtained solid was precipitated in methanol and filtered to reach compound 12 as a red solid (30 mg, 94%). 1H NMR (300 MHz, C6D6) δ= 7.79 (d, J= 8.5 Hz, 4H), 7.26 (d, J= 8.5 Hz, 4H), 1.18 (s, 18H). 19F NMR (282 MHz, C6D6) δ= −137.88 (d, J= 15.2 Hz), −154.87 (d, J= 15.2 Hz). 13C{1H}{19F} NMR (126 MHz, C6D6) δ= 153.2, 143.9, 139.6, 131.9, 126.1, 121.4, 120.7, 111.9, 107.3, 87.5, 31.1, 30.2. FT-IR (cm-1) 3086, 3040, 2955, 2906, 2865, 2191, 1668, 1502, 1444, 1407, 1363, 1267, 1076, 834, 659, 561. HRMS found: 634.1880. Calculated for C38H26F8, 634.1907. 9,10-Bis[penta(4-tert-butyl(phenyl))phenyl] octafluoroanthracene (6b). To a 30 mL dry microwave reactor vial provided with a magnetic stir bar, 2,3,4,5-tetrakis[4-(tert- butyl)phenyl]cyclopenta-2,4-dien-1-one 5 (0.098 mmol, 59 mg, 3.1 equiv.) and 12 (0.031 mmol, 20 mg, 1 equiv.) were added. The reaction was carried out in a microwave reactor at 280 ºC and 600 rpm for 40 Experimental section 148 minutes, repeating these conditions for a total of three sequential times. The resulting crude was purified through silica gel chromatography column using as eluent a gradient from hexane to CS2 affording 6b as a yellow solid (12 mg, 21%). 1H NMR (500 MHz, CDCl3) δ= 6.81 (m, 12H), 6.76 (d, J= 8.0 Hz, 4H), 6.70 – 6.61 (m, 20H), 6.16 (d, J= 8.0 Hz, 4H), 1.11 (s, 18H), 1.09 (s, 36H), 1.08 (s, 36H). 19F NMR (471 MHz, CDCl3) δ= −137.53 (d, J= 15.9 Hz), −160.43 (d, J= 15.9 Hz). 13C{1H}{19F} NMR (126 MHz, CDCl3) δ= 148.1, 147.7, 141.9, 140.3, 140.3, 137.9, 137.7, 136.4, 131.8, 131.6, 131.3, 130.0, 129.3, 123.5, 123.3, 122.4, 118.6, 34.2, 31.3, 31.3, 31.2, 29.9. FT-IR (cm-1) 3088, 3033, 2961, 2904, 2967, 1675, 1493, 1454, 1362, 1270, 1072, 1015, 832, 570. HRMS found:1795.0137. Calculated for C126H130F8, 1795.0045. Helically arranged nanographene 13. In a 100 mL dry flask, under argon atmosphere and provided with a magnetic stir bar, polyphenylene 6b (0.006 mmol, 10 mg, 1 equiv.) was solved in 15 mL of anhydrous DCM. The mixture was cooled to 0 ºC and stirred for five minutes after DDQ (0.245 mmol, 56 mg, 44 equiv.) addition. Then, under Ar bubbling, TfOH was added dropwise (2.290 mmol, 344 mg, 412 equiv.) and the reaction was carried out at 0 ºC maintaining the Ar bubbling for 40 minutes. After that time a HNaCO3 saturated solution was added to the reaction flask and the crude was extracted with DCM and washed twice with a HNaCO3 saturated solution (30 mL) and twice with brine (30 mL). From this step the organic phase was dried with MgSO4, and the solvent was removed under reduced pressure. The crude was manipulated and purified in light absence through a silica gel column in hexane:toluene 50:1, affording 13 as a yellow solid (6 mg, 63%). 1H NMR (500 MHz, CDCl3) δ= 9.22 (s, 4H), 9.18 (d, J= 1.7 Hz, 4H), 8.99 (d, J= 1.7 Hz, 4H), 8.71 (d, J= 2.0 Hz, 1H), 7.54 (d, J= 8.9 Hz, 4H), 6.98 (dd, J= 8.9, 2.0 Hz, 4H), 1.83 (s, 18H), 1.78 (s, 36H), 1.20 (s, 36H). 19F NMR (471 MHz, CDCl3) δ= −137.41 (d, J= 12.85 Hz), -157.37 (d, J= 12.85 Hz). 13C{1H}{19F} NMR (126 MHz, CDCl3) δ= 149.6, 149.6, 135.0, 132.3, 131.4, 130.6, 130.3, 130.0, 129.7, 127.4, 126.0, 123.5, 123.0, 122.9, 122.0, 121.8, 120.5, 119.5, 119.2, 119.0, 118.5, 35.9, 35.8, 34.8, 32.2, 32.1, 31.0. FT-IR (cm-1) 2958, 2924, 2875, 2859, 1464, 1375. HRMS found: 1778.8759. Calculated for C126H114F8, 1778.8793. Chapter 2 1,4-Bis[(4-(tert-butyl)phenyl)ethynyl]tetrafluoro benzene (15a). To a dry 10 mL microwave reactor vial under argon atmosphere and provided with a magnetic stir bar, CuI (0.06 mmol, 12 mg, 0.4 equiv.), Pd(dppf)Cl2 (0.03 mmol, 27 mg, 0.2 equiv.), 1,4-dibromotetrafluorobenzene 14 (0.16 mmol, 50 mg, 1 equiv.), 4-(tert- butyl)phenylacetylene 3a (0.41 mmol, 64 mg, 2.5 equiv.), 1.7 mL of anhydrous previously deoxygenated THF and 0.4 mL of distilled DIPA were added. The reaction was carried out for 3 hours at 120 ºC and 600 rpm in a microwave reactor. The Experimental section 149 resulting crude was extracted with DCM and washed with NH4Cl (50 mL), water (50 mL) and brine (50 mL). The organic phase was dried with MgSO4, filtered through a silica gel plug column using dichloromethane as eluent and the solvent was removed under reduced pressure. The product was dissolved in the minimum amount of CHCl3 and then methanol was added to obtain a precipitate that was filtered to reach 15a as white solid (59 mg, 79%). 1H NMR (300 MHz, Chloroform-d) δ= 7.54 (d, J= 8.5 Hz, 4H), 7.42 (d, J= 8.5 Hz, 4H), 1.34 (s, 18H). 19F NMR (471 MHz, Chloroform-d) δ= −138.00 (s). 13C{1H) NMR (126 MHz, Chloroform-d) δ= 153.3, 146.6 (dm, 1JC-F= 243.5 Hz), 131.9, 125.7, 118.8, 104.9 (m) 103.4, 74.3, 35.1, 31.2. FT-IR (cm-1) 3030, 2959, 2904, 2867, 2229, 2210, 1605, 1522, 1512, 1487, 980, 949, 835. HRMS: Calculated for C30H26F4: 462.1971, found: 462.1958. 1,4-bis[(4-methoxyphenyl)ethynyl]tetrafluoro benzene (15b). To a dry 10 mL microwave reactor vial under argon atmosphere and provided with a magnetic stir bar, CuI (0.06 mmol, 12 mg, 0.4 equiv.), Pd(dppf)Cl2 (0.03 mmol, 27 mg, 0.2 equiv.), 1,4-dibromotetra fluorobenzene 14 (0.16 mmol, 50 mg, 1 equiv.), 4- (methoxy)phenylacetylene 3b (0.41 mmol, 54 mg, 2.5 equiv.), 1.7 mL of anhydrous previously deoxygenated THF and 0.4 mL of distilled DIPA were added. The reaction was carried out for 3 hours at 120 ºC and 600 rpm in a microwave reactor. The resulting crude was extracted with DCM and washed twice with NH4Cl (50 mL), twice with water (50 mL) and twice with brine (50 mL). The organic phase was dried with MgSO4, filtered through a silica gel plug column using dichloromethane as eluent and solvent was removed under reduced pressure. The resulting solid was dissolved in the minimum amount of CHCl3 and then methanol was added to obtain a precipitate that was filtered to reach 15b as a brown solid (42 mg, 64%). 1H NMR (300 MHz, Chloroform-d) δ= 7.54 (d, J= 8.8 Hz, 4H), 6.91 (d, J= 8.8 Hz, 4H), 3.85 (s, 6H). 19F NMR (282 MHz, Chloroform-d) δ= −138.3 (s). 13C{1H} NMR (126 MHz, Chloroform-d) δ= 160.9, 146.6 (dm, 1JC-F= 261.9 Hz), 133.8, 114.3, 113.9, 104.8 (m), 103.3, 73.9, 55.5. FT-IR (cm-1) 3022, 3002, 2970, 2937, 2909, 2840, 2228, 2210, 1601, 1518, 1481, 1246, 1172, 1024, 965, 830. HRMS: Calculated for C24H14F4O2: 410.0930, found: 410.0910. 1,4-Bis(phenylethynyl)tetrafluorobenzene (15c). To a dry 10 mL microwave reactor vial under argon atmosphere and provided with a magnetic stir bar, CuI (0.13 mmol, 20 mg, 0.4 equiv.), Pd(dppf)Cl2 (0.06 mmol, 50 mg, 0.2 equiv.), 1,4-dibromotetrafluorobenzene 14 (0.32 mmol, 100 mg, 1 equiv.), phenylacetylene 3c (0.81 mmol, 80 mg, 2.5 equiv.), 3.3 mL of anhydrous previously deoxygenated THF and 0.9 mL of distilled DIPA were added. The reaction was carried out for 3 hours at 120 ºC and 600 rpm in a microwave reactor. The crude was extracted with DCM and washed twice with NH4Cl (40 mL), twice with water (40 mL) and twice with brine (40 mL). The organic phase was dried with MgSO4, filtered through a silica gel plug column using dichloromethane as eluent Experimental section 150 and solvent was removed under reduced pressure. The product was dissolved in the minimum amount of CHCl3 and then methanol was added to obtain a precipitate that was filtered to reach to 15c as beige solid (80 mg, 70%). 1H NMR (300 MHz, Chloroform-d) δ= 7.64 – 7.57 (m, 4H), 7.46 – 7.36 (m, 6H). 19F NMR (471 MHz, Chloroform-d) δ= −137.70 (s). 13C{1H} (126 MHz, Chloroform-d) δ= 146.7 (dm, 1JC- F= 254.2 Hz), 132.1, 129.9, 128.7, 121.8, 104.9, 103.2, 74.8. FT-IR (cm-1) 3060, 2232, 2210, 1484, 1443, 975. HRMS: Calculated for C22H10F4: 350.0719, found: 350.0703. 1,4-Bis[penta(4-tert-butyl(phenyl))phenyl]tetra fluorobenzene 16a. To a 10 mL dry microwave reactor vial, provided with a magnetic stir bar, 2,3,4,5-tetrakis[4-(tert-butyl)phenyl]cyclopenta- 2,4-dien-1-one 5 (0.34 mmol, 200 mg, 3.1 equiv.) and 15a (0.09 mmol, 45 mg, 1 equiv.) were added. The reaction was carried out in a microwave reactor at 300 ºC and 600 rpm for 30 minutes, repeating these conditions for a total of four sequential times and removing carbon monoxide under vacuum between each repetition. The resulting crude was dissolved in the minimum amount of CHCl3, and methanol was added to obtain a precipitate that was filtered, affording 16a as a white solid (110 mg, 70%). 1H NMR (500 MHz, Chloroform-d) δ= 7.18 (d, J = 8.1 Hz, 4H), 6.83 (d, J = 8.3 Hz, 4H), 6.78 (d, J = 8.6 Hz, 12H), 6.68 (d, J = 8.3 Hz, 4H), 6.61 (d, J = 8.1 Hz, 4H), 6.59 – 6.48 (m, 12H), 1.25 (s, 36H), 1.08 – 1.07 (m, 54H). 19F NMR (282 MHz, Chloroform-d) δ= −137.48 (s). 13C{1H} NMR (126 MHz, Chloroform-d) δ= 148.5, 147.7, 147.6, 143.1 (dm, 1JC-F =256.2 Hz), 142.6, 141.9, 141.3, 137.7, 137.6, 137.0, 132.39, 131.2, 131.0, 128.9, 125.6, 123.6, 123.2, 122.8, 119.0 (m), 34.5, 34.2, 31.6, 31.3. FT-IR (cm-1) 3089, 3039, 2962, 2903, 2869, 1899, 1773, 1657, 1609, 1512, 1469, 1391, 1362, 1268, 1020, 978, 831. HRMS: Calculated for C118H130F4=1623.0109, found=1623.0033. 1,4-Bis[(2-[4-methoxyphenyl)-3,4,5,6-tetra(4- tert-butyl(phenyl))]phenyl]tetrafluorobenzene 16b. To a 10 mL dry microwave reactor vial, provided with a magnetic stir bar, 2,3,4,5- tetrakis[4-(tert-butyl)phenyl] cyclopenta-2,4- dien-1-one 5 (0.23 mmol, 140 mg, 3.1 equiv.) and 15b (0.07 mmol, 30 mg, 1 equiv.) were added. The reaction was carried out in a microwave reactor at 300 ºC and 600 rpm for 30 minutes, repeating these conditions for a total of three times and removing carbon monoxide under vacuum between each repetition. The resulting crude was dissolved in the minimum amount of CHCl3 and methanol was added to obtain a precipitate that was filtered, affording 16b as a nude-colored solid (64 mg, 56%). 1H NMR (500 MHz, Chloroform-d) δ= 7.00 (d, J= 8.4 Hz, 2H), 6.81 (d, J= 8.5 Hz, 2H), 6.77 – 6.73 (m, 4H), 6.66 – 6.48 (m, 12H), 3.73 (s, 3H), 1.23 (s, 9H), 1.10 (s, 9H), 1.07 (s, 9H), 1.06 (s, 9H). 19F NMR (471 MHz, Chloroform-d) δ= −138.32 (s). 13C{1H} NMR (126 MHz, Chloroform-d) δ= 157.7, 148.5, 147.7, 147.7, 147.6, 143.3 Experimental section 151 (dm, 1JC-F = 246.7 Hz), 142.8, 141.9, 141.4, 141.2, 141.1, 137.7, 137.6, 137.5, 137.0, 132.4, 131.8, 131.1, 131.0, 130.9, 130.5, 125.7, 123.5, 123.4, 123.2, 123.1, 119.3 (m), 112.4, 55.3, 34.5, 34.2, 34.2, 31.5, 31.3, 31.3. FT-IR (cm-1) 3087, 3032, 2959, 2901, 2865, 1899, 1779, 1742, 1661, 1611, 1513, 1467, 1245, 977, 832. HRMS: Calculated for C112H118F4O2: 1570.9068, found: 1570.9106. 1,4-Bis[2-phenyl-3,4,5,6-tetra(4-tert- butyl(phenyl))phenyl]tetrafluorobenzene 16c. To a 10 mL dry microwave reactor vial, provided with a magnetic stir bar, 2,3,4,5-tetrakis[4-(tert- butyl)phenyl]cyclopenta-2,4-dien-1-one 5 (0.35 mmol, 210 mg, 3.1 equiv.) and 15c (0.11 mmol, 48 mg, 1 equiv.) were added. The reaction was carried out in a microwave reactor at 280 ºC and 600 rpm for 30 minutes, repeating these conditions for a total of three times and removing carbon monoxide under vacuum between each repetition. The resulting crude dissolved in the minimum amount of CHCl3, and methanol was added to obtain a precipitate that was filtered, affording 16c as a pale brown solid (141 mg, 82%). 1H NMR (500 MHz, Chloroform- d) δ= 6.96 (d, J= 8.3 Hz, 4H), 6.88 (t, J= 7.2 Hz, 2H), 6.82 – 6.63 (m, 24H), 6.59 – 6.57 (m, 8H), 6.49 (d, J= 8.3 Hz, 4H), 1.24 (s, 18H), 1.10 – 1.02 (m, 54H). 19F NMR (282 MHz, Chloroform-d) δ= −138.89 (s). 13C{1H}{19F} NMR (126 MHz, Chloroform-d) δ= 148.25, 147.8, 147.7, 147.4, 143.3, 142.7, 141.9, 141.8, 141.4, 140.7, 139.7, 137.6, 137.5, 137.3, 137.0, 131.0, 130.9, 130.8, 130.2, 130.1, 126.7, 125.7, 125.1, 123.8, 123.3, 123.1, 123.0, 119.5, 34.5, 34.2, 34.1, 31.5, 31.3. FT-IR (cm-1) 3090, 3051, 3028, 2960, 2905, 2865, 1896, 1779, 1661, 1603, 1513, 1462, 1391, 1361, 1269, 1022, 976, 832. HRMS: Calculated for C110H114F4: 1510.8857, found: 1510.8841. Helically arranged nanographene 17a. To a 100 mL dry flask, under argon atmosphere and provided with a magnetic stir bar, 16a (0.01 mmol, 20 mg, 1 equiv.), 37 mL of DCM and DDQ (0.54 mmol, 123 mg, 44 equiv.) were added. The reaction mixture was cooled to 0 ºC and stirred for 5 minutes. Then, under Ar bubbling, TfOH was added dropwise (5.07 mmol, 760 mg, 412 equiv.) and the reaction was carried out at 0 ºC maintaining the Ar bubbling for 25 minutes. After that time, a HNaCO3 saturated solution was added to the reaction flask and the crude was extracted with DCM and washed twice with a HNaCO3 saturated solution (40 mL) and twice with brine (40 mL). The organic phase was dried with MgSO4, and the solvent was removed under reduced pressure. The crude was purified by aluminum oxide column chromatography using pentane as eluent, affording 17a as a yellow solid (13 mg, 80%). 1H NMR (300 MHz, Chloroform-d) δ= 9.16 (s, 4H), 9.11 (d, J= 1.6 Hz, 4H), 9.01 (d, J= 1.6 Hz, 4H), 8.81 (d, J= 1.9 Hz, 4H), 8.26 (d, J= 8.8 Hz, 4H), 7.64 (dd, J= 8.8, 1.9 Hz, 4H), 1.82 (s, 18H), 1.76 (s, 36H), 1.39 (s, 36H). 19F NMR Experimental section 152 (282 MHz, Chloroform-d) δ= −137.26 (s). 13C{1H} NMR (126 MHz, Chloroform-d) δ= 150.5, 149.9, 149.8, 147.0 (d, 1JC-F= 241.1 Hz), 132.3, 130.7, 130.2, 130.0, 128.3, 127.5, 126.3, 124.2 (m), 124.2, 123.3, 123.2, 122.6, 122.1, 122.0, 120.2, 119.3, 119.2, 118.8, 118.5, 35.9, 35.8, 35.1, 32.2, 32.1, 31.3, 29.9. FT-IR (cm-1) 3082, 2952, 2921, 2852, 1608, 1588, 1479, 1463, 1374, 1246, 981, 871. HRMS: Calculated for C118H114F4: 1606.8857, found: 1606.8866. Helically arranged nanographene 17b. To a 100 mL dry flask, under argon atmosphere and provided with a magnetic stir bar, 16b (0.01 mmol, 20 mg, 1 equiv.), 38 mL of DCM and DDQ (0.56 mmol, 127 mg, 44 equiv.) were added. The reaction mixture was cooled to 0 ºC and stirred for 5 minutes. Then, under Ar bubbling, TfOH was added dropwise (5.24 mmol, 790 mg, 412 equiv.) and the reaction was carried out at 0 ºC maintaining the Ar bubbling for 25 minutes. After that time, a HNaCO3 saturated solution was added to the reaction flask and the crude was extracted with DCM and washed twice with a HNaCO3 saturated solution (40 mL) and twice with brine (40 mL). The organic phase was dried with MgSO4, and the solvent was removed under reduced pressure. The crude was filtered through silica gel using DCM as eluent and solvent was removed under reduced pressure affording 17b as a green-yellow solid (11 mg, 70%, 1:1 anti/syn). 1H NMR (500 MHz, Chloroform-d) δ= 9.20 – 9.09 (m, 8H), 9.02 (s, 1H), 8.99 (s, 1H), 8.93 (s, 1H), 8.91 (s, 1H), 8.82 (d, J= 2.0 Hz, 1H), 8.80 (d, J= 2.0 Hz, 1H), 8.30 – 8.23 (m, 4H), 8.21 (d, J= 9.2 Hz, 1H), 8.15 (d, J= 8.8 Hz, 1H), 7.62 (dd, J= 8.8, 2.0 Hz, 1H), 7.54 (dd, J= 8.8, 2.0 Hz, 1H), 7.15 (dd, J= 9.2, 2.6 Hz, 1H), 7.11 (dd, J= 9.2, 2.6 Hz, 1H), 3.95 (s, 3H), 3.93 (s, 3H), 1.81 (s, 18H), 1.77 – 1.74 (m, 36H), 1.39 – 1.38 (m, 18H). 19F NMR (471 MHz, Chloroform-d) δ= −137.16 (s), −137.17 (s). 13C{1H}{19F} NMR (126 MHz, Chloroform-d) δ= 159.0, 159.0, 150.5, 150.5, 149.9, 149.8, 147.0, 134.4, 134.4, 132.3, 132.3, 130.7, 130.6, 130.6, 130.5, 130.2, 130.1, 130.0, 129.9, 129.8, 129.7, 129.2, 128.3, 128.2, 127.4, 127.4, 126.5, 126.5, 126.4, 124.6, 124.4, 124.2, 124.1, 123.4, 123.4, 123.2, 123.2, 122.9, 122.9, 122.7, 122.7, 122.6, 122.1, 122.0, 121.9, 120.1, 120.1, 119.8, 119.7, 119.4, 119.3, 119.2, 118.8, 118.7, 118.6, 118.5, 118.3, 119.9, 111.6, 108.2, 107.7, 55.6, 55.5, 35.9, 35.8, 35.1, 32.1, 32.1, 31.4, 31.3. FT-IR (cm-1) 3075, 2960, 2909, 2865, 1785, 1722, 1610, 1589, 1458, 1372, 1248, 1044, 979, 870. HRMS: Calculated for C112H102F4O2: 1554.7816, found: 1554.7840. Experimental section 153 Helically arranged nanographene 17c. To a 100 mL dry flask, under argon atmosphere and provided with a magnetic stir bar, 16c (0.01 mmol, 20 mg, 1 equiv.), 40 mL of DCM and DDQ (0.58 mmol, 132 mg, 44 equiv.) were added, the reaction mixture was cooled to 0 ºC and stirred for 5 minutes. Then, under Ar bubbling, TfOH was added dropwise (5.45 mmol, 820 mg, 412 equiv.) and the reaction was carried out at 0 ºC maintaining the Ar bubbling for 25 minutes. After that time, a HNaCO3 saturated solution was added to the reaction flask and the crude was extracted with DCM and washed twice with a HNaCO3 saturated solution (40 mL) and twice with brine (40 mL). The organic phase was dried with MgSO4, and the solvent was removed under reduced pressure. The crude was purified by silica gel chromatography column performed using a gradient from hexane:DCM (12:1) to hexane:DCM (6:1), solvent was removed under reduced pressure affording 17c as a yellow solid (12 mg, 77%, 1:1 anti/syn). 1H NMR (500 MHz, Chloroform-d) δ= 9.17 – 9.15 (m, 4H), 9.13 (s, 2H), 9.11 (s, 2H), 9.02 – 8.99 (m, 4H), 8.87 − 8.84 (m, 2H), 8.81 (d, J= 1.8 Hz, 1H), 8.80 (d, J= 1.8 Hz, 1H), 8.32 (d, J= 8.4 Hz, 1H), 8.29 (d, J= 8.4 Hz, 1H), 8.24 (d, J= 8.8 Hz, 1H), 8.15 (d, J= 8.8 Hz, 1H), 7.68 – 7.64 (m, 2H), 7.61 (dd, J= 8.8, 1.8 Hz, 1H), 7.54 – 7.50 (m, 3H), 1.81 (s, 18H), 1.76 − 1.74 (m, 36H), 1.38 (s, 9H), 1.37 (s, 9H). 19F NMR (471 MHz, Chloroform-d) δ= −137.35 (s), −137.37 (s). 13C{1H}{19F} NMR (126 MHz, Chloroform-d) δ= 150.6, 150.5, 150.0, 150.0, 150.0, 149.9, 149.9, 146.9, 132.7, 132.6, 132.3, 132.2, 130.8, 130.7, 130.6, 130.6, 130.6, 130.3, 130.2, 130.1, 130.1, 130.0, 130.0, 129.9, 128.3, 128.2, 127.8, 127.7, 127.6, 127.5, 126.4, 126.3, 124.4, 124.3, 124.2, 124.1, 124.0, 123.5, 123.4, 123.3, 123.3, 123.2, 123.1, 123.1, 122.6, 122.1, 122.0, 121.9, 121.9, 120.2, 120.1, 119.5, 119.5, 119.4, 119.3, 119.3, 119.2, 119.0, 118.9, 118.7, 118.6, 118.5, 35.9, 35.8, 35.1, 32.2, 32.1, 31.4, 31.3. FT-IR (cm-1) 3081, 2959, 2905, 2866, 1607, 1588, 1476, 1458, 1372, 1243, 979, 870. HRMS: Calculated for C110H98F4: 1494.7605, found: 1494.7620. Nanographene 17c’. To a 100 mL dry flask, under argon atmosphere and provided with a magnetic stir bar, 16c (0.01 mmol, 10 mg, 1 equiv.), 20 mL of DCM and DDQ (0.29 mmol, 66 mg, 44 equiv.) were added. The reaction mixture was cooled to −78 ºC and stirred for 5 minutes. Then, under Ar bubbling, TfOH was added dropwise (2.72 mmol, 410 mg, 412 equiv.) and the reaction was carried out at −78 ºC maintaining the Ar bubbling for 3 hours. After that time, a HNaCO3 saturated solution was added to the reaction flask and the crude was extracted with DCM and washed twice with a HNaCO3 saturated solution (20 mL) and twice with brine (20 mL). The organic phase was dried with MgSO4 and filtered through silica gel using DCM as eluent. Solvent was removed under reduced pressure affording 17c’ as a yellow solid (7 mg, 75%). 1H NMR (700 MHz, 1,1,2,2- Experimental section 154 Tetrachloroethane-d2) δ= 9.16 (s, 1H), 9.15 (s, 1H), 9.01 (s, 1H), 8.99 (s, 1H), 8.77 (d, J= 2.1 Hz, 1H), 8.69 (d, J= 2.1 Hz, 1H), 7.86 (d, J= 8.6 Hz, 1H), 7.64 (d, J= 8.8 Hz, 1H), 7.36 (d, J= 7.4 Hz, 2H), 7.33 (dd, J= 8.6, 2.1 Hz, 1H), 7.23 (t, J= 7.4 Hz, 1H), 7.18 (t, J= 7.4 Hz, 2H), 7.15 (dd, J= 8.8, 2.1 Hz, 1H), 1.77 (s, 9H), 1.76 (s, 9H), 1.60 (s, 9H), 1.46 (s, 9H). 19F NMR (282 MHz, 1,1,2,2-Tetrachloroethane-d2) δ= −135.52. 13C{1H} (176 MHz, 1,1,2,2-Tetrachloroethane-d2) δ= 149.6, 149.4, 149.4, 149.2, 144.1 (dm, 1JC-F = 236.1 Hz), 142.8, 142.3, 140.1, 131.0, 130.9, 130.6, 130.4, 130.3, 129.7, 129.3, 128.3, 128.1, 128.0, 127.8, 127.6, 127.4, 126.9, 125.4, 125.0, 123.7, 123.5, 122.7 (m), 122.0, 121.9, 121.3, 120.6, 120.2, 119.0, 118.9, 118.7, 118.4, 35.4, 34.9, 34.7, 31.8, 31.7, 31.4, 31.2. Calculated for C110H102F4: 1498.7918, found: 1498.7570. Chapter 3 (Pentahelicen-2,13-diyldiethyne-2,1-diyl)bis[tris(1-methyl ethyl)silane] (19a). A pressure tube was charged with dichlorohelicene 18a (300 mg, 0.865 mmol, 1.0 equiv.), Pd2(dba)3 (16 mg, 0.017 mmol, 2 mol%), XPhos (33 mg, 0.0688 mmol, 8 mol%), potassium carbonate (713 mg, 5.160 mmol, 6.0 equiv.), flushed with nitrogen, and degassed (by three freeze-pump-thaw cycles) DMF (15 mL) was added. Then (triisopropylsilyl)acetylene (770 µL, 3.44 mmol, 4.0 equiv.) was added and the reaction mixture was stirred at 120 °C for 16 h. The solvent was evaporated under reduced pressure and the residue was purified by flash chromatography on silica gel (hexane to hexane-ethyl acetate 20:1) to afford diyne 19a (418 mg, 76%) as an amorphous solid. 1H NMR (400 MHz, CDCl3) δ= 8.64 (bd, J= 1.3 Hz, 2H), 7.83 – 7.90 (m, 8H), 7.53 (dd, J= 8.3, 1.5 Hz, 2H), 1.02 – 1.09 (m, 42H). 13C NMR (101 MHz, CDCl3) δ= 133.2, 132.6, 132.1, 130.4, 128.9, 127.7, 127.5, 127.4, 127.1, 126.6, 120.0, 107.7, 90.4, 18.9, 18.8, 11.5. IR, CHCl3 (cm−1): 3064, 3049, 2959, 2944, 2923, 2891, 2865, 2152, 1606, 1502, 1463, 1438, 1383, 1367, 1301, 1189, 1072, 996, 883, 849, 678, 660, 634. APCI MS: 639 ([M+H]+). HR APCI MS: Calculated for C44H55Si2: 639.3842, found: 639.3835. (Heptahelicen-2,17-diyldiethyne-2,1-diyl)bis[tris(1-methyl ethyl)silane] (19b). A pressure tube was charged with helicene 18b (450 mg, 1.011 mmol, 1.0 equiv.), Pd2(dba)3 (18 mg, 0.0202 mmol, 2 mol%), XPhos (38 mg, 0.0807 mmol, 8 mol%), potassium carbonate (837 mg, 6.060 mmol, 6.0 equiv.), flushed with nitrogen, and degassed (by three freeze-pump-thaw cycles) DMF (20 mL) was added. Then (triisopropylsilyl)acetylene (910 µL, 4.041 mmol, 4.0 equiv.) was added and the reaction mixture was stirred at 120 °C for 16 h. The solvent was evaporated under reduced pressure and the residue was purified by flash chromatography on silica gel (hexane to hexane-ethyl acetate 20:1) to afford diyne 19b (530 mg, 71%) as an amorphous solid. 1H NMR (400 MHz, CDCl3) δ= 8.04 (s, 2H), 8.00 (d, J= 8.2 Hz, 2H), 7.87 (d, J= 8.2 Hz, 2H), 7.73 (d, J= 8.5 Hz, 2H), 7.50 (d, J= 8.5 Hz, 2H), 7.30 – Experimental section 155 7.33 (m, 2H), 7.27 (d, J= 8.3 Hz, 2H), 6.99 (dd, J= 8.2, 1.6 Hz, 2H), 1.09 – 1.16 (m, 42H). 13C NMR (101 MHz, CDCl3) δ= 132.2, 131.53, 131.46, 129.1, 128.8, 128.1, 127.8, 127.7, 127.4, 127.1, 127.0, 126.8, 126.7, 125.0, 118.9, 107.6, 89.0, 18.9, 11.5. IR, CHCl3 (cm−1): 3055, 2959, 2944, 2925, 2892, 2865, 2152, 1606, 1518, 1495, 1463, 1428, 1390, 1383, 1366, 1308, 1191, 1073, 996, 883, 847, 679, 660, 641. APCI MS: 739 ([M+H]+). HR ESI MS: Calculated for C52H59Si2: 739.4155, found: 739.4148. 2,13-Bis[(4-(tert-butyl)phenyl)ethynyl]pentahelicene (20a). A 25 mL dry round bottom flask provided with a magnetic stir bar was charged with helicene 19a (100 mg, 0.156 mmol, 1 equiv.), Pd(PPh3)2Cl2 (12 mg, 0.018 mmol, 0.1 equiv.) and CuI (125 mg, 0.656 mmol, 4 equiv.). The flask was evacuated and backfilled with dry Ag three times to exclude moisture and air. Dry THF (1 mL), freshly distilled DIPA (1.5 mL) and 4-tert- butyliodobenzene (86 mg, 0.329 mmol, 2.0 equiv) were added to the flask and stirred for 5 minutes before adding tetrabutylammoniumfloride (TBAF 1.0 M in THF, 0.5 mL, 0.5 mmol, 3.0 equiv) in one portion. The mixture was stirred at 30 °C for 16 hours. Silica was poured into the flask and solvent was removed under reduced pressure to adsorb the mixture which was then purified by silica gel column chromatography with hexane:DCM 10:1 affording the product 20a as a white solid (83 mg, 86%). 1H NMR (300 MHz, CDCl3) δ= 8.79 (s, 2H), 7.95 – 7.85 (m, 8H), 7.65 (dd, J= 8.3, 1.4 Hz, 2H), 7.37 (d, J= 8.0 Hz, 4H), 7.30 (d, J= 8.5 Hz, 4H), 1.30 (d, J= 0.9 Hz, 18H). 13C NMR (75 MHz, CDCl3) δ= 151.4, 132.8, 132.2, 132.1, 131.5, 130.4, 129.2, 128.1, 127.7, 127.5, 127.2, 126.7, 125.3, 120.6, 120.0, 89.7, 89.4, 34.9, 31.3. FT-IR (cm−1): 3044, 2958, 2863, 2149, 1907, 1793, 1698, 1607, 1517, 1463, 1363, 1266, 846, 727, 674, 561. HR-MS MALDI-ToF: Calculated for C46H38: 590.2974, found: 590.2994. 2,17-Bis[(4-(tert-butyl)phenyl)ethynyl]heptahelicene 20b. A 25 mL dry Schlenk provided with a magnetic stir bar was charged with helicene 19b (100 mg, 0.135 mmol, 1 equiv.), Pd(PPh3)2Cl2 (10 mg, 0.013 mmol, 0.1 equiv.) and CuI (100 mg, 0.541 mmol, 4 equiv.). The Schlenk was evacuated and backfilled with dry Ag two times to exclude moisture and air. Dry THF (5 mL), freshly distilled DIPA (1.4 mL) and 4-tert-butyliodobenzene (140 mg, 0.541 mmol, 4.0 equiv) were added to the Schlenk and stirred for 5 minutes before adding tetrabutylammoniumfloride (TBAF 1.0 M in THF, 0.41 mL, 0.41 mmol, 3.0 equiv) in one portion. The mixture was stirred at 30 °C for 20 hours. The crude was poured into a flask and silica was added, solvent was removed under reduced pressure to adsorb the mixture which was then purified by silica gel column chromatography with hexane:DCM 100:20 affording the product 20b as a white solid (63 mg, 67%). 1H NMR (300 MHz, CDCl3) δ= 8.07 (s, 2H), 8.07 (d, J= 8.2 Hz, 2H), 8.00 (d, J= 8.2 Hz, 2H), 7.79 (d, J= 8.5 Hz, 2H), 7.54 (d, J= 8.5 Hz, 2H), 7.44 – 7.38 (m, 10H), 7.32 (d, J= 8.2 Hz, 2H), 7.04 (dd, J= 8.2, 1.6 Hz, 2H), 1.37 (s, 18H). 13C NMR (126 MHz, Experimental section 156 CDCl3) δ= 151.3, 132.2, 131.7, 131.5, 131.3, 129.1, 128.8, 127.8, 127.6, 127.5, 127.4, 127.1, 127.04, 127.01, 126.9, 125.5, 125.1, 120.9, 118.9, 89.7, 88.8, 35.0, 31.4. FT- IR (cm−1): 3048, 2959, 2923, 2858, 1907, 1732, 1605, 1503, 1462, 1363, 1264, 1102, 1017, 834, 549. HR-MS MALDI-ToF: Calculated for C54H42: 690.3287, found: 690.3267. 2,13-Bis[penta(4-tert-butyl(phenyl))phenyl]pentahelicene 21a To a 30 mL microwave vial provided with a magnetic stir bar was added 20a (40 mg, 0.068 mmol, 1.0 equiv) and tetra- 2,3,4,5-tetrakis[4-(1,1-dimethylethyl)phenyl]2,4-cyclopen tadien-1-one 5 (128 mg, 0.210 mmol, 3.1 equiv). The vial was placed in the microwave reactor and heated to 280 °C with a hold time of 40 minutes and stirring at 300 rpm, same conditions were repeated for 6 times. After cooling, methanol was added, and the crude was sonicated to obtain a suspension. The solid was filtered to reach to compound 21a as a brown solid (33 mg, 28%). 1H NMR (300 MHz, CDCl3) δ= 8.42 (s, 2H), 7.55 (d, J= 8.5 Hz, 2H), 7.46 (d, J= 8.5 Hz, 2H), 7.34 (s, 2H), 7.23 (s, 2H), 7.05 – 6.57 (m, 32H), 6.51 (d, J= 8.7 Hz, 2H), 6.11 (dd, J= 8.1, 2.0 Hz, 2H), 5.99 (d, J= 9.1 Hz, 2H), 5.92 (d, J= 8.3 Hz, 2H), 5.84 (d, J= 8.3 Hz, 2H), 1.24 (s, 18H), 1.14 (s, 18H), 1.09 (s, 18H), 1.01 (s, 18H), 0.57 (s, 18H). 13C NMR (126 MHz, CDCl3) δ= 148.1, 147.3, 147.2, 147.1, 146.7, 141.3, 141.0, 140.9, 140.2, 139.7, 138.5, 138.4, 138.2, 138.0, 137.9, 137.9, 135.7, 133.1, 132.0, 131.6, 131.2, 130.7, 130.4, 130.1, 129.4, 128.9, 127.2, 126.1, 126.0, 125.7, 124.9, 123.8, 123.0, 122.8, 122.2, 122.1, 34.4, 34.0, 34.0, 33.9, 33.3, 31.5, 31.3, 31.2, 31.1, 30.8. FT-IR (cm−1): 3088, 3053, 2962, 2905, 2870, 1910, 1746, 1615, 1514, 1462, 1393, 1361, 1270, 1120, 1020, 832, 575. HR-MS MALDI-ToF: Calculated for C134H142: 1751.1111, found: 1751.1151. 2,17-Bis[penta(4-tert-butyl(phenyl))phenyl]pentahelicene 21b. To a 10 mL microwave vial provided with a magnetic stir bar was added 20b (30 mg, 0.043 mmol, 1.0 equiv) and tetra- 2,3,4,5-tetrakis[4-(1,1-dimethylethyl)phenyl]2,4-cyclopenta dien-1-one 5 (87 mg, 0.14 mmol, 3.1 equiv). The vial was placed in the microwave reactor and heated to 280 °C with a hold time of 45 minutes and stirring at 600 rpm. After cooling, methanol was added, and the crude was sonicated to obtain a suspension. The solid was filtered to reach to compound 21b as a brown solid (36 mg, 45%). 1H NMR (500 MHz, CDCl3) δ= 7.89 (s, 2H), 7.81 (d, J= 8.2 Hz, 2H), 7.72 (d, J= 8.2 Hz, 2H), 7.59 z(d, J= 8.6 Hz, 2H), 7.55 (d, J= 8.6 Hz, 2H), 7.13 (d, J= 8.2 Hz, 2H), 6.88 – 6.50 (m, 34H), 6.47 (d, J= 8.1 Hz, 2H), 5.96 (d, J= 8.1 Hz, 2H), 5.85 (d, J= 8.0 Hz, 2H), 5.65 (s, 4H), 1.14 (s, 18H), 1.12 (s, 18H), 1.09 (s, 18H), 1.06 (s, 18H), 0.43 (s, 18H).13C NMR (126 MHz, CDCl3) δ= 147.5, 147.4, 147.3, 147.2, 146.2, 140.6, 140.3, 140.3, 139.1, 138.7, 138.4, 138.2, 137.6, 137.2, 136.9, 132.6, 132.2, 132.0, 131.7, 131.3, 131.2, 131.1, 131.0, 130.4, 129.5, 129.2, 129.0, 128.4, 127.9, Experimental section 157 127.4, 126.2, 125.7, 125.4, 124.7, 124.6, 123.2, 123.1, 123.0, 122.9, 122.3, 122.1, 34.2, 34.2, 34.1, 34.1, 33.3, 31.6, 31.4, 31.4, 31.3, 30.7. FT-IR (cm−1): 3051, 2961, 2904, 2866, 1899, 1695, 1610, 1510, 1461, 1361, 1270, 831, 572. HR-MS MALDI- ToF: Calculated for C142H146: 1851.1424, found: 1851.1429. [9]Helical bilayer nanographene ([9]HBNG). To a 100 mL flask with magnetic stir bar was added polyphenylene 21a (20 mg, 0.011 mmol, 1 equiv.). The flask was purged and backfilled with argon and dry dicloromethane (50 mL) was added. A glass Pasteur pipette was attached to an argon inlet and dipped into the solution. Argon was bubbled continuously through the solution, while stirring, for the duration of the reaction. The mixture was cooled to 0 °C in an ice water bath and 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ) (360 mg, 1.586 mmol, 140 equiv.) was added in one portion. The mixture was stirred for 5 minutes and trifluoromethanesulfonic acid (1.20 mL, 2.052 g, 13.67 mmol, 1200 equiv) was added dropwise. The mixture was stirred at 0 °C for 20 minutes. After that time a HNaCO3 saturated solution was added to the reaction flask and the crude was extracted with DCM and washed twice with a HNaCO3 saturated solution (50 mL) and twice with brine (50 mL). From this step the organic phases were dried with MgSO4 and filtered, the solvent was removed under reduced pressure. The mixture was purified by silica gel column in hexane:DCM 10:1, affording compound [9]HBNG as a yellow solid (12 mg, 61%). 1H NMR (500 MHz, CDCl3) δ= 9.18 (s, 2H), 9.13 (s, 2H), 9.06 (s, 2H), 8.87 (s, 2H), 8.75 (s, 2H), 8.71 (s, 2H), 8.39 (s, 2H), 8.21 (s, 2H), 8.20 (s, 2H), 8.14 (s, 2H), 8.13 (s, 2H), 7.96 (d, J= 8.1 Hz, 2H), 7.86 (d, J= 8.1 Hz, 2H), 7.61 (s, 2H), 1.93 (s, 18H), 1.86 (s, 18H), 1.82 (s, 18H), 1.43 (s, 18H), 1.21 (s, 18H). 13C NMR (126 MHz, CDCl3) δ= 147.9, 147.7, 147.4, 146.8, 146.5, 130.4, 130.3, 130.3, 130.2, 129.9, 129.4, 129.3, 129.0, 128.9, 128.8, 128.7, 127.9, 127.9, 127.4, 126.8, 126.6, 125.7, 123.7, 123.6, 123.5, 123.4, 122.9, 122.6, 122.3, 119.8, 119.8, 119.6, 119.5, 119.3, 119.2, 119.0, 118.9, 118.7, 118.3, 118.0, 117.3, 117.2, 117.2, 116.7, 35.9, 35.8, 35.6, 35.2, 34.6, 32.6, 32.3, 32.3, 32.0, 31.6. FT-IR (cm−1): 3102, 3046, 2953, 2906, 2868, 1745, 1606, 1575, 1478, 1361, 1261, 1201, 867. HR-MS MALDI-ToF: Calculated for C134H118: 1726.9233, found: 1726.9267. [11]Helical bilayer nanographene ([11]HBNG). To a 250 mL flask with magnetic stir bar was added polyphenylene 21b (55 mg, 0.094 mmol, 1 equiv.). The flask was purged and backfilled with argon and dry dicloromethane (90 mL) was added. A glass Pasteur pipette was attached to an argon inlet and dipped into the solution. Argon was bubbled continuously through the solution, while stirring, for the duration of the reaction. The mixture was cooled to 0 °C in an ice water bath and DDQ (320 mg, 1.409 mmol, 44 equiv) was added in one portion. The mixture was stirred for 5 minutes and trifluoromethanesulfonic acid (1.16 mL, 1.984 g, 13.22 mmol, 141 equiv) Experimental section 158 was added dropwise. The mixture was stirred at 0 °C for 1 hour, after that time a HNaCO3 saturated solution was added to the reaction flask and the crude was extracted with DCM and washed twice with a HNaCO3 saturated solution (90 mL) and twice with brine (90 mL). From this step the organic phases were dried with MgSO4 and filtered, the solvent was removed under reduced pressure. The mixture was purified by silica gel column in hexane:DCM 80:20, affording compound [11]HBNG as a yellow solid (42 mg, 75%). 1H NMR (700 MHz, CDCl3) δ = 9.47 (s, 4H), 9.46 (d, J= 7.4 Hz, 4H), 9.38 (d, J= 7.4 Hz, 4H), 9.12 (s, 2H), 8.95 (s, 2H), 8.44 (s, 2H), 8.20 (s, 2H), 7.35 (s, 2H), 6.93 (d, J= 1.6 Hz, 2H), 6.90 (d, J= 7.5 Hz, 2H), 6.11 – 6.08 (m, 4H), 5.32 (d, J= 7.9 Hz, 2H), 1.96 (s, 18H), 1.95 (s, 18H), 1.94 (s, 18H), 1.87 (s, 18H), 0.89 (s, 18H). 13C NMR (176 MHz, CDCl3) δ = 149.3, 149.2, 149.1, 146.4, 131.6, 131.1, 131.0, 130.9, 130.8, 130.7, 130.7, 130.6, 130.4, 130.4, 129.1, 128.1, 127.5, 126.9, 126.2, 125.9, 125.2, 125.2, 124.9, 124.9, 124.6, 124.5, 124.2, 123.9, 123.7, 123.0, 121.3, 121.1, 120.6, 120.5, 120.5, 120.2, 119.9, 119.9, 119.8, 119.2, 119.1, 119.0, 119.0, 118.7, 118.5, 118.3, 118.0, 36.0, 36.0, 36.0, 34.4, 32.6, 32.3, 31.3. FT-IR (cm−1): 3077, 3045, 2955, 2906, 2870, 1605, 1572, 1479, 1369, 1257, 870. HR-MS MALDI-ToF: Calculated for C142H122: 1826.9547, found: 1826.9566. Chapter 4 7,7'-dibromo-1,1'-binaphthyl-2,2'-diol 23. A 10 mL tube provided with a magnetic stir bar was charged with 7-bromo-2-naphtol 22 (500 mg, 2.241 mmol, 1 equiv.), Cu-TMDA(OH)(Cl) (50 mg, 0.108 mmol, 0.05 equiv.) and dissolved in DCM (2.5 mL). The mixture was sonicated for 5 minutes, the stirred at 25 ºC for 16 h under air. The crude was extracted in ethyl acetate washing with water (30 mL) and brine (30 mL). The organic phase was dried with Na2SO4, and filtered, and the solvent was removed under reduced pressure. The resulting solid was dissolved in the minimum amount of DCM, then, hexane was added, and the mixture was sonicated until precipitation. The mixture was filtered to obtained compound 23 as a white solid (414 mg, 83%). 1H NMR (300 MHz, CDCl3) δ= 7.97 (d, J= 9.0 Hz, 1H), 7.78 (d, J= 8.7 Hz, 1H), 7.48 (dd, J= 8.7, 1.9 Hz, 1H), 7.40 (d, J= 9.0 Hz, 1H), 7.23 (d, J= 1.9 Hz, 1H), 5.01 (s, 1H).171 7,7'-bis[(4-(tert-butyl)phenyl)ethynyl]-1,1'-binaphthyl- 2,2'-diol 24. To a dry 30 mL microwave reactor vial under argon atmosphere and provided with a magnetic stir bar, CuI (120 mg, 0.630 mmol, 0.4 equiv.), Pd(PPh3)2Cl2 (360 mg, 0.315 mmol, 0.2 equiv.), 7,7'-dibromo-1,1'-binaphthyl-2,2'-diol 23 (700 mg, 1.58 mmol, 1 equiv.), 4-(tert-butyl)phenylacetylene 3 (740 mg, 4.73 mmol, 3 equiv.), 5 mL of anhydrous previously deoxygenated DMF and 3 mL of distilled triethylamine were added. The reaction was carried out for 3 hours at 85 ºC and 600 rpm in a microwave 171 Y. Nojima, M. Hasegawa, N. Hara, Y. Imai, Y. Mazakia, Chem. Commun. 2019, 55, 2749. Experimental section 159 reactor. The resulting crude was extracted with DCM and washed two times with HCl 1M (50 mL), water (50 mL) and brine (50 mL). The organic phase was dried with MgSO4, and solvent was removed under reduced pressure. The crude was purified by silica gel chromatography column in hexane/ethyl acetate 4:1. Solvent was removed, and the product was recrystallized by adding the minimum amount of DCM and hexane, obtaining compound 24 as white needles (908 mg, 96%). 1H NMR (500 MHz, CDCl3) δ= 7.99 (d, J= 8.9 Hz, 1H), 7.88 (d, J= 8.4 Hz, 1H), 7.50 (dd, J= 8.4, 1.6 Hz, 1H), 7.41 (d, J= 8.9 Hz, 1H), 7.38 (d, J= 9.1 Hz, 2H), 7.31 – 7.29 (m, 3H), 5.03 (s, 1H), 1.29 (s, 9H). 13C NMR (126 MHz, CDCl3) δ= 153.5, 151.8, 133.4, 131.6, 131.5, 129.0, 128.7, 127.3, 125.4, 122.9, 120.0, 118.6, 110.5, 90.7, 89.3, 34.9, 31.3. FT-IR (cm−1): 3530, 3333, 3068, 2961, 2864, 1899, 1617, 1505, 1442, 1330, 1192, 1153, 836, 560. HRMS: Calculated for C44H38O2: 598.2872, found: 598.2851. M*-7,7'-Bis[penta(4-tert-butyl(phenyl)) phenyl]-1,1'-binaphthyl-2,2'-diol 25. To a 30 mL microwave vial provided with a magnetic stir bar was added 24 (200 mg, 0.334 mmol, 1.0 equiv.) and tetra-2,3,4,5-tetrakis[4-(1,1-di methylethyl)phenyl]2,4-cyclopentadien-1-one 5 (610 mg, 1.002 mmol, 3 equiv.). The vial was placed in the microwave reactor and heated to 280 °C with a hold time of 45 minutes and stirring at 600 rpm. After cooling, the crude was dissolved in the minimum amount of DCM, and methanol was added to obtain a suspension. The solid was filtered washing with methanol affording compound 25 as a white solid (547 mg, 93%). 1H NMR (500 MHz, CDCl3) δ= 7.64 (d, J= 8.8 Hz, 1H), 7.42 (d, J= 8.4 Hz, 1H), 7.04 – 6.99 (m, 2H), 6.94 (s, 1H), 6.85 – 6.71 (m, 10H), 6.67 – 6.58 (m, 8H), 6.50 (d, J= 8.3 Hz, 2H), 3.74 (s, 1H), 1.09 (s, 9H), 1.07 (s, 27H), 1.05 (s, 9H). 13C NMR (126 MHz, CDCl3) δ= 152.3, 147.6, 147.53, 147.48, 147.3, 147.3, 141.0, 140.8, 140.7, 140.64, 140.59, 140.2, 139.4, 138.1, 138.01, 137.98, 137.8, 137.7, 132.4, 131.5, 131.2, 131.16, 131.1, 131.0, 130.7, 130.4, 130.2, 128.2, 127.4, 126.8, 126.4, 124.2, 123.5, 123.4, 123.2, 123.1, 123.0, 122.98, 122.94, 122.9, 117.5, 111.8, 34.23, 34.19, 34.18, 34.15, 34.11, 31.42, 31.34, 31.32, 31.28. FT-IR (cm−1): 3536, 3055, 2961, 2905, 2864, 1895, 1617, 1508, 1363, 1268, 832, 571. HRMS: Calculated for C132H142O2: 1759.1010, found: 1759.1040. Experimental section 160 M*-7,7'-bis(penta(4-tert-butyl(phenyl))phenyl) -2'-hydroxy-[1,1'-binaphthalen]-2-yl (1R)-10- camphorsulfonate 26a-b. A 10 mL tube provided with a magnetic stir bar was charged with BINOL- polyarene 25 (100 mg, 0.057 mmol, 1 equiv.) and (1R)-10-camphor sulphonyl chloride (142 mg, 0.568 mmol, 10 equiv.). The tube was evacuated and backfilled with dry Ag two times to exclude moisture and air. Dry DCM (2 mL) and anhydrous pyridine (0.05 mL) were added at 0 ºC. The mixture was stirred at 0 ºC for 5 h and warmed to 20 ºC for 16 h. The crude was extracted in chloroform washing with NH4Cl (50 mL) and brine (50 mL). The organic phase was dried with Mg2SO4, and filtered, and the solvent was removed under reduced pressure. The crude was purified by silica gel chromatography column with toluene, affording diastereomer 26a (30 mg, 27%) and diastereomer 26b (28 mg, 25%). M-7,7'-bis(penta(4-tert-butyl(phenyl))phenyl)-2'-hydroxy-[1,1'-binaphthalen]- 2-yl (1R)-10-camphorsulfonate 26a 1H NMR (500 MHz, CDCl3) δ= 7.79 (d, J= 9.0 Hz, 1H), 7.49 (d, J= 9.1 Hz, 1H), 7.47 (d, J= 8.3 Hz, 1H), 7.42 (d, J= 8.9 Hz, 1H), 7.28 (d, J= 8.6 Hz, 1H), 7.16 (d, J= 8.4 Hz, 1H), 7.15 (s, 1H), 7.02 (s, 1H), 6.96 (s, 1H), 6.91 – 6.88 (m, 2H), 6.84 – 6.70 (m, 24H), 6.65 – 6.54 (m, 12H), 6.52 (dd, J= 8.7 Hz, J= 2.4 Hz, 1H), 6.40 (m, 3H), 3.53 (s, 1H), 2.47 (d, J= 15.1 Hz, 1H), 2.22 – 2.09 (m, 2H), 1.90 – 1.83 (m, 2H), 1.80 (d, J= 15.1 Hz, 1H), 1.72 (d, J= 18.4 Hz, 1H), 1.11 (s, 9H), 1.10 (s, 9H), 1.09 – 1.07 (m, 54H), 1.06 (s, 9H), 1.01 (s, 9H), 0.84 (s, 3H), 0.58 (s, 3H). 13C NMR (126 MHz, CDCl3) δ= 213.2, 150.9, 148.0, 147.9, 147.6, 147.6, 147.5, 147.4, 147.3, 147.2, 146.3, 141.3, 141.1, 141.03, 140.99, 140.8, 140.7, 140.6, 140.35, 140.28, 140.2, 140.1, 139.9, 138.5, 138.2, 138.2, 138.0, 137.9, 137.9, 137.8, 137.7, 137.3, 132.6, 132.5, 132.1, 131.7, 131.5, 131.3, 131.2, 131.0, 130.8, 130.6, 129.9, 128.5, 127.7, 127.5, 126.9, 126.6, 126.3, 124.2, 123.5, 123.5, 123.1, 122.9, 122.2, 117.4, 114.3, 57.8, 48.0, 47.6, 42.9, 42.3, 34.29, 34.27, 34.17, 34.15, 34.13, 34.11, 34.07, 31.49, 31.45, 31.35, 31.33, 31.31, 26.8, 25.1, 19.9, 19.5. FT-IR (cm−1): 3534, 3060, 2964, 2864, 1911, 1735, 1512, 1363, 1160, 830, 571. HRMS: Calculated for C142H156O5S: 1973.1673, found: 1973.1664. P-7,7'-bis(penta(4-tert-butyl(phenyl))phenyl) -2'-hydroxy-[1,1'-binaphthalen]-2- yl (1R)-10-camphorsulfonate 26b 1H NMR (500 MHz, CDCl3) δ= 7.79 (d, J= 9.0 Hz, 1H), 7.49 (d, J= 8.9 Hz, 1H), 7.50 – 7.44 (m, 2H), 7.24 (s, 1H), 7.15 – 7.13 (m, 2H), 7.07 (s, 1H), 6.98 (s, 1H), 6.89 – 6.36 (m, 40H), 6.41 (d, J= 8.2 Hz, 2H), 6.37 (d, J= 8.3 Hz, 1H), 3.54 (s, 1H), 2.49 (d, J= 14.9 Hz, 1H), 2.21 – 2.12 (m, 1H), 2.08 – 2.03 (m, 1H), 1.93 – 1.86 (m, 2H), 1.85 – 1.80 (m, 1H), 1.72 (d, J= 18 Hz, 1H), 1.12 (s, 9H), 1.11 (s, 9H), 1.09 (s, 9H), 1.08 (s, 9H), 1.07 (s, 9H), 1.07 (s, 9H), 1.07 (s, 9H), 1.06 (s, 9H), 1.05 (s, 9H), 1.00 (s, 9H), 0.81 (s, 3H), 0.60 (s, 3H). 13C NMR (126 MHz, CDCl3) δ= 213.2, 151.0, 147.8, 147.8, 147.7, 147.6, 147.5, 147.4, 147.3, 147.2, 146.4, Experimental section 161 141.3, 141.1, 141.0, 140.8, 140.7, 140.5, 140.3, 140.3, 140.2, 140.2, 140.1, 139.9, 138.6, 138.3, 138.2, 138.0, 138.0, 137.96, 137.94, 137.8, 137.7, 137.3, 132.7, 132.6, 132.1, 131.8, 131.2, 131.1, 131.0, 130.9, 130.6, 129.8, 129.6, 128.7, 127.7, 127.6, 126.9, 126.5, 126.1, 124.5, 124.0, 123.7, 123.5, 123.4, 123.2, 123.1, 123.0, 122.8, 122.2, 117.7, 114.7, 58.0, 47.9, 47.5, 43.3, 42.4, 34.31, 34.26, 34.18, 34.16, 34.14, 34.12, 34.06, 31.50, 31.45, 31.36, 31.34, 31.32, 31.30, 26.8, 25.5, 19.9, 19.7. M-7,7'-Bis[penta(4-tert-butyl(phenyl)) phenyl]-1,1'-binaphthyl-2,2'-diol (M- 25). A 10 mL tube provided with a magnetic stir bar was charged with 26a (30 mg, 0.015 mmol, 1 equiv.) and 1 mL of THF. Then NaOH (10 mg, 0.243 mmol, 16 equiv.) and NBu4Br (49 mg, 0.152 mmol, 10 equiv.) were dissolved in 0.2 mL of distilled water and added to the reaction mixture. The mixture was stirred at 50 ºC for 16 h. The crude was extracted in DCM washing with NH4Cl (30 mL) and brine (30 mL). The organic phase was dried with Mg2SO4, and filtered, the solvent was removed under reduced pressure. The crude was purified by silica gel chromatography column with toluene, affording diastereomer M-25 as a white solid (25 mg, 92%). The same procedure was followed from 26b to obtain P-25 as a white solid (27 mg 99%). M*- oxa[9]Helical bilayer nanographene (oxa[9]HBNG). To a 250 mL flask with magnetic stir bar was added polyphenylene 25 (60 mg, 0.034 mmol, 1 equiv.). Dry dichloromethane (60 mL) was added. A glass Pasteur pipette was attached to an argon inlet and dipped into the solution. Argon was bubbled continuously through the solution, while stirring, for the duration of the reaction. The mixture was cooled to 0 °C in an ice water bath and DDQ (464 mg, 2.040 mmol, 60 equiv.) was added in one portion. The mixture was stirred for 5 minutes and trifluoromethanesulfonic acid (0.6 mL, 1.021 g, 6.183 mmol, 200 equiv.) was added dropwise. The mixture was stirred at 0 °C for 25 minutes, after that time a HNaCO3 saturated solution was added to the reaction flask and the crude was extracted with DCM and washed twice with a HNaCO3 saturated solution (100 mL) and twice with brine (100 mL). From this step the organic phases were dried with MgSO4 and filtered, the solvent was removed under reduced pressure. The mixture was purified by silica gel column in hexane:DCM 70:30, affording compound oxa[9]HBNG as a yellow solid (36 mg, 61%). 1H NMR (500 MHz, CDCl3) δ= 9.54 (s, 1H), 9.35 (s, 1H), 9.34 (s, 1H), 9.29 (s, 1H), 9.24 (s, 1H), 9.00 (s, 1H), 8.83 (d, J= 8.4 Hz, 1H), 8.58 (d, J= 8.4 Hz, 1H), 8.35 (s, 1H), 7.98 (s, 1H), 7.62 (s, 1H), 7.39 (s, 1H), 7.15 (s, 1H), 1.98 (s, 9H), 1.91 (s, 9H), 1.70 (s, 9H), 0.97 (s, 9H), 0.62 (s, 9H). 13C NMR (126 MHz, CDCl3) δ= 154.6, 148.8, 148.4, 147.8, 147.4, 144.9, 130.6, 130.3, 130.1, 130.0, 129.96, 129.3, 129.2, Experimental section 162 128.55, 128.5, 128.3, 128.1, 127.8, 126.6, 124.1, 124.0, 123.9, 123.6, 123.2, 121.8, 121.3, 121.0, 120.9, 120.7, 120.6, 120.3, 119.8, 119.5, 119.3, 119.1, 118.8, 118.7, 118.6, 118.4, 118.0, 117.9, 117.3, 115.9, 112.5, 35.9, 35.8, 35.5, 34.7, 34.4, 32.32, 32.30, 32.2, 31.4, 30.7. FT-IR (cm−1): 3086, 2951, 2903, 2867, 1856, 1741, 1601, 1577, 1477, 1462, 1360, 1259, 1232, 941, 966. HRMS: Calculated for C132H116O: 1716.9026, found: 1716.8948. M-oxa[9]helical bilayer nanographene (M-oxa[9]HBNG). To a 50 mL flask with magnetic stir bar was added polyphenylene M-25 (20 mg, 0.011 mmol, 1 equiv.). Dry dichloromethane (20 mL) was added. A glass Pasteur pipette was attached to an argon inlet and dipped into the solution. Argon was bubbled continuously through the solution, while stirring, for the duration of the reaction. The mixture was cooled to −30 °C in an ice water bath and DDQ (62 mg, 0.273 mmol, 24 equiv.) was added in one portion. The mixture was stirred for 5 minutes and trifluoromethanesulfonic acid (0.1 mL, 1.136 mmol, 100 equiv.) was added dropwise. The mixture was stirred at −30 °C for 50 minutes, after that time a HNaCO3 saturated solution was added to the reaction flask and the crude was extracted with DCM and washed twice with a HNaCO3 saturated solution (100 mL) and twice with brine (100 mL). From this step the organic phases were dried with MgSO4 and filtered, the solvent was removed under reduced pressure. The mixture was purified by silica gel preparative in CS2, affording compound M-oxa[9]HBNG as a yellow solid (10 mg, 51%). P-oxa[9]helical bilayer nanographene (P-oxa[9]HBNG). To a 50 mL flask with magnetic stir bar was added polyphenylene P-25 (25 mg, 0.014 mmol, 1 equiv.). Dry dichloromethane (25 mL) was added. A glass Pasteur pipette was attached to an argon inlet and dipped into the solution. Argon was bubbled continuously through the solution, while stirring, for the duration of the reaction. The mixture was cooled to −30 °C in an ice water bath and DDQ (77 mg, 0.341 mmol, 24 equiv.) was added in one portion. The mixture was stirred for 5 minutes and trifluoromethanesulfonic acid (0.12 mL, 1.420 mmol, 100 equiv.) was added dropwise. The mixture was stirred at −30 °C for 45 minutes, after that time a HNaCO3 saturated solution was added to the reaction flask and the crude was extracted with DCM and washed twice with a HNaCO3 saturated solution (100 mL) and twice with brine (100 mL). From this step the organic phases were dried with MgSO4 and filtered, the solvent was removed under reduced pressure. The mixture was purified by silica gel preparative in CS2, affording compound P-oxa[9]HBNG as a yellow solid (10 mg, 40%). Appendix 1 Appendix 1. Chapter 1 165 APPENDIX 1. Chapter 1 Crystallographic Data Single crystals suitable for X-ray diffraction analysis were obtained for both isomers of compound 7, for compound syn 8, and also for compound syn 10. Crystals of both isomers of 7 and compound syn 10 are not stable out of the mother liquor and quickly lose crystallinity. The structure refinements of syn 7, anti 7 and syn 10 present high wR2 values due to the high degree of disorder that many of the atoms in the asymmetric unit display. As it is usual in the crystal structures of nanographene molecules, solvent molecules in the interstices are extremely disordered, as well as some of the peripheral tBu substituents. The coordinates of many of the atoms in these tBu fragments had to be refined with the aid of geometrical restraints and/or two alternative sets of positions, and for some of the terminal methyl groups the positions found in the electron density maps could not account for 100% of the expected electron density. Structural details of syn 7 (CCDC 2154765) Figure AP1.1. Labelled asymmetric unit of compound syn 7 (hydrogen atoms and solvent molecules have been omitted for clarity). In the disordered tBu groups only one of the alternative sets of position is represented. Table AP1.1. Sample and crystal data for syn 7. Chemical formula (C126H136)·3.75(H2O) Formula weight 1717.90 g/mol Temperature 250(2) K Wavelength 0.71073 Å Crystal habit clear colorless needle Crystal system triclinic Space group P -1 a = 12.093 (2) Å α = 98.008(5)° Appendix 1. Chapter 1 166 Unit cell dimensions b = 16.919(2) Å β = 90.350(5)° c = 30.799(3) Å γ = 108.296(4)° Volume 5916.4(12) Å3 Density (calculated) 0.964 g/cm3 Z 2 Table AP1.2. Data collection and structure refinement for syn 7. Theta range for data collection 1.28 to 25.24° Index ranges -14<=h<=14, -20<=k<=20, -36<=l<=36 Reflections collected 61800 Independent reflections 20875 [R(int) = 0.1106] Coverage of independent reflections 97.5% Absorption correction multi-scan Max. and min. transmission 0.9970 and 0.9820 Refinement method Full-matrix least-squares on F2 Refinement program SHELXL-2014/7 (Sheldrick, 2014) Function minimized Σ w(Fo 2 - Fc 2)2 Data / restraints / parameters 20875 / 987 / 1134 Goodness-of-fit on F2 1.117 Final R indices 7699 data; I>2σ(I) R1 = 0.1887, wR2 = 0.4676 all data R1 = 0.3301 wR2 = 0.5390 Weighting scheme w=1/[σ2(Fo 2)+(0.2000P)2+21.700P] where P=(Fo 2+2Fc 2)/3 Structural details of anti 7 (CCDC 2154766) Figure AP1.2. Labelled asymmetric unit of compound anti 7 (hydrogen atoms and solvent molecules have been omitted for clarity). Appendix 1. Chapter 1 167 Table AP1.3. Sample and crystal data for anti 7. Chemical formula (C126H136)·0.25 (CH3OH)·3.75 (H2O) Formula weight 1725.91 g/mol Temperature 150(2) K Wavelength 0.71073 Å Crystal habit colorless prismatic Crystal system triclinic Space group P -1 Unit cell dimensions a = 17.2815(16) Å α = 65.267(4)° b = 19.0056(17) Å β = 72.847(4)° c = 19.6699(18) Å γ = 88.649(4)° Volume 5569.6(9) Å3 Density (calculated) 1.029 g/cm3 Z 2 Table AP1.4. Data collection and structure refinement for anti 7. Theta range for data collection 1.19 to 25.03° Index ranges -20<=h<=20, -22<=k<=22, -23<=l<=23 Reflections collected 105318 Independent reflections 19632 [R(int) = 0.1960] Coverage of independent reflections 99.7% Absorption correction multi-scan Max. and min. transmission 0.9990 and 1.0000 Refinement method Full-matrix least-squares on F2 Refinement program SHELXL-2014/7 (Sheldrick, 2014) Function minimized Σ w(Fo 2 - Fc 2)2 Data / restraints / parameters 9632 / 1434 / 1179 Goodness-of-fit on F2 1.017 Final R indices 7130 data; I>2σ(I) R1 = 0.2129, wR2 = 0.4765 all data R1 = 0.3573, wR2 = 0.5926 Weighting scheme w=1/[σ2(Fo 2)+(0.3200P)2+19.7000P] where P=(Fo 2+2Fc 2)/3 Appendix 1. Chapter 1 168 Structural details of syn 8 (CCDC 2209634) Figure AP1.3. Labelled asymmetric unit of compound syn 8 (hydrogen atoms and solvent molecules have been omitted for clarity). In the disordered tBu groups only one of the alternative sets of position is represented. Table AP1.5. Sample and crystal data for syn 8. Chemical formula C126H130 Formula weight 1644.29 g/mol Temperature 100(2) K Wavelength 1.54184 Å Crystal habit pale yellow plate Crystal system triclinic Space group P -1 Unit cell dimensions a = 16.8928(3) Å α = 71.5968(18)° b = 18.4600(4) Å β = 65.0663(18)° c = 19.9255(4)Å γ = 62.9236(19)° Volume 4957.6(2) Å3 Density (calculated) 1.102 g/cm3 Z 2 Table AP1.6. Data collection and structure refinement for syn 8. Theta range for data collection 2.72 to 68.28° Index ranges -19<=h<=20, -22<=k<=21, -20<=l<=24 Reflections collected 63593 Independent reflections 18134 [R(int) = 0.0350] Coverage of independent reflections 99.8% Absorption correction multi-scan Max. and min. transmission 1.000 and 0.863 Refinement method Full-matrix least-squares on F2 Refinement program SHELXL-2019/1 (Sheldrick, 2019) Appendix 1. Chapter 1 169 Function minimized Σ w(Fo 2 - Fc 2)2 Data / restraints / parameters 18134 / 69 / 1318 Goodness-of-fit on F2 1.034 Final R indices 13621 data; I>2σ(I) R1 = 0.0510, wR2 = 0.1320 all data R1 = 0.0717, wR2 = 0.1442 Weighting scheme w=1/[σ2(Fo 2)+(0.0754P)2+1.3997P] where P=(Fo 2+2Fc 2)/3 Structural details of syn 10 (CCDC 2209635) Figure AP1.4. Labelled asymmetric unit of compound syn 10 (hydrogen atoms and solvent molecules have been omitted for clarity). In the disordered tBu groups only one of the alternative sets of position is represented. Table AP1.7. Sample and crystal data for syn 10. Chemical formula (C126H120)·2.5 (C2H4Cl2)·2 (H2O) Formula weight 956.80 g/mol Temperature 150(2) K Wavelength 0.71073 Å Crystal habit clear intense yellow prismatic Crystal system triclinic Space group P -1 Unit cell dimensions a = 14.1128(7) Å α = 70.4452(19)° b = 17.4044(8) Å β = 84.838(2)° c = 25.6992(10) Å γ = 84.805(2)° Volume 5911.9(5) Å3 Density (calculated) 1.075 g/cm3 Z 4 Appendix 1. Chapter 1 170 Table AP1.8. Data collection and structure refinement for syn 10. Theta range for data collection 1.69 to 25.03° Index ranges -16<=h<=16, -20<=k<=20, -30<=l<=30 Reflections collected 74424 Independent reflections 20878 [R(int) = 0.0953] Coverage of independent reflections 99.9% Absorption correction multi-scan Max. and min. transmission 0.9860 and 0.9400 Refinement method Full-matrix least-squares on F2 Refinement program SHELXL-2019/1 (Sheldrick, 2019) Function minimized Σ w(Fo 2 - Fc 2)2 Data / restraints / parameters 20878 / 1501 / 1469 Goodness-of-fit on F2 1.040 Final R indices 10247 data; I>2σ(I) R1 = 0.1189, wR2 = 0.3283 all data R1 = 0.2137, wR2 = 0.4036 Weighting scheme w=1/[σ2(Fo 2)+(0.200P)2+19.7000P] where P=(Fo 2+2Fc 2)/3 Theoretical Calculations All the calculations reported in this work were obtained with the GAUSSIAN 09 suite of programs.172 Electron correlation was partially taken into account using the B3LYP173 functional in conjuction with the D3 dispersion correction suggested by Grimme et al.174 and the double- quality plus polarization functions def2-SVP175 basis set for all atoms. Reactants and products were characterized by frequency calculations,176 and have positive definite Hessian matrices. Transition states (TSs) show only one negative eigenvalue in their diagonalized force constant matrices, and their associated eigenvectors were confirmed to correspond to the motion along the reaction coordinate. Connections between reactants and products were confirmed 172 Gaussian 09, Revision D.01, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman, and D. J. Fox, Gaussian, Inc., Wallingford CT, 2016. 173 (a) A. D. Becke, J. Chem. Phys. 1993, 98, 5648; b) C. Lee, W. Yang, R. G. Parr, Phys. Rev. B 1998, 37, 785; c) S. H. Vosko, L. Wilk, M. Nusair, Can. J. Phys. 1980, 58, 1200. 174 S. Grimme, J. Antony, S. Ehrlich, H. Krieg, H. J. Chem. Phys. 2010, 132, 154104. 175 F. Weigend, R. Alhrichs, Phys. Chem. Chem. Phys. 2005, 7, 3297. 176 J. W. McIver, A. K. Komornicki, J. Am. Chem. Soc. 1972, 94, 2625. Appendix 1. Chapter 1 171 using the Intrinsic Reaction Coordinate (IRC) method.177 Solvent effects were considered by using the Polarizable Continuum Model (PCM).178 This level is denoted PCM(solvent)-B3LYP-D3/def2-SVP. Calculations of the absorption spectrum were accomplished using time-dependent density functional theory (TD-DFT)179 at the B3LYP-D3/def2-SVP level using the optimized geometries. The assignment of the excitation energies to the experimental bands was performed on the basis of the energy values and oscillator strengths. The B3LYP Hamiltonian was chosen because it was proven to provide reasonable UV-vis spectra for a variety of chromophores.180 The aromaticity of the considered species has been assessed by the computation of the NICS181 values using the gauge invariant atomic orbital (GIAO) method,182 at the B3LYP183 level using the def2-SVP basis set, with the optimized PCM(solvent)- B3LYP-D3/def2-SVP geometries. Non-covalent interactions in compound 10M were assessed by means of the NCIPlot method.184 Cartesian coordinates (in Å) and total energies (in a.u., ZPVE included) of all the stationary points discussed in the text. All calculations have been performed at the PCM(solvent)-B3LYP-D3/def2-SVP level. 177 C. González, H. B. Schlegel, J. Phys. Chem. 1990, 94, 5523. 178 (a) S. Miertuš, E. Scrocco, J. Tomasi, Chem. Phys. 1981, 55, 117; (b) J. L. Pascual-Ahuir, E. Silla, I. Tuñón, J. Comp. Chem. 1994, 15, 1127; (c) V. Barone, M. Cossi, J. Phys. Chem. A, 1998, 102, 1995. 179 (a) M. E. Casida, Recent Developments and Applications of Modern Density Functional Theory; Elsevier: Amsterdam, 1996; Vol. 4; (b) M. E. Casida, D. P. Chong, Recent Advances in Density Functional Methods; World Scientific: Singapore, 1995; Vol. 1, p 155. 180 For a review, see: A. Dreuw, M. Head-Gordon, M. Chem. Rev. 2005, 105, 4009. 181 Chen, Z.; Wannere, C. S.; Corminboeuf, C.; Puchta, R.; Schleyer, P. v. R. Chem. Rev. 2005, 105, 3842. 182 Wolinski, K.; Hilton, J. F.; Pulay, P. J. Am. Chem. Soc. 1990, 112, 8251. 183 (a) Becke, A. D. J. Chem. Phys. 1993, 98, 5648; (b) Lee, C.; Yang, W.; Parr, R. G. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785; (c) Vosko, S. H.; Wilk, L.; Nusair, M. Can. J. Phys. 1980, 58, 1200. 184 Johnson, E. R.; Keinan, S.; Mori-Sánchez, P.; Contreras-García, J.; Cohen, A. J.; Yang, W. J. Am. Chem. Soc. 2010, 132, 6498. Appendix 1. Chapter 1 172 Figure AP1.5. Radical-cation mechanism towards spironanographenes Appendix 2 Appendix 2. Chapter 2 175 APPENDIX 2. Chapter 2 Crystallographic Data Single crystals suitable for X-ray diffraction analysis were obtained for compound 17a. The molecule presents chirality, although the crystals obtained from the racemic mixture are racemic. These crystals are not stable out of the mother liquor and quickly lose crystallinity. The structure refinement presents a high wR2 value due to the high degree of disorder that many of the atoms in the asymmetric unit display. The solvent molecules (dichloroethane and water) in the interstices are extremely disordered, as well as the peripheral tBu substituents. The coordinates of the atoms in the tBu fragments had to be refined with the aid of geometrical restraints, and for some of the terminal methyl groups (C60-C62, C104-C106, C108-C110, C112-C114, C116-C118) these positions only account for 50% of the corresponding electron density. Figure AP2.1. Labelled asymmetric unit of compound 17a (solvent molecules are omitted). Structural details of 17a (CCDC 2070237) Table AP2.1. Sample and crystal data for 17a. Chemical formula (C118H114F4)·(C2H2Cl2)0.5(H2O)4.5 Formula weight 1737.72 Temperature 200(2) K Wavelength 0.71073 Å Crystal size 0.105 x 0.221 x 0.261 mm Crystal habit clear orange needle Crystal system monoclinic Space group C2/c Unit cell dimensions a = 35.348(2) Å α = 90° Appendix 2. Chapter 2 176 b = 31.054(1) Å β = 114.553(2)° c = 22.447(1) Å γ = 90° Volume 22412(2) Å3 Z 8 Table AP2.2. Data collection and structure refinement for 17a. Theta range for data collection 1.14 to 23.96° Index ranges -33<=h<=40, -34<=k<=28, -25<=l<=16 Reflections collected 43262 Independent reflections 16808 [R(int) = 0.0678] Coverage of independent reflections 96% Absorption correction multi-scan Max. and min. transmission 0.9910 and 0.9780 Structure solution technique direct methods Structure solution program SHELXS-97 (Sheldrick 2008) Refinement method Full-matrix least-squares on F2 Refinement program SHELXL-2014/7 (Sheldrick, 2014) Function minimized Σ w(Fo 2 - Fc 2)2 Data / restraints / parameters 16808/33/1057 Goodness-of-fit on F2 1.123 Final R indices 8046 data; I>2σ(I) R1 = 0.1564, wR2 = 0.4257 all data R1 = 0.2562 wR2 = 0.5056 Weighting scheme w=1/[σ2(Fo 2)+(0.3200P)2+19.700P] where P=(Fo 2+2Fc 2)/3 Appendix 2. Chapter 2 177 Chiral HPLC resolution Analytical HPLC analysis was collected on an Agilent Technologies 1260 Infinity HPLC system using Regis Pirkle-type covalently modified silica column (4.6 × 250 mm) Whelk-O2 (R,R). Injection volume was 0.7 μL, flow rate was 1 mL/min and eluent was 98:1:1 hexane:isopropanol:tetrahydrofuran. Peaks were confirmed by comparison of their UV-Vis spectra to originate from the enantiomers. Figure AP2.2. HPLC chromatogram of racemic 17a recorded at 360 nm. Appendix 3 Appendix 3. Chapter 3 181 APPENDIX 3. Chapter 3 Crystallographic data Single crystals suitable for X ray diffraction analysis were obtained from the respective racemic mixtures containing P and M isomers of compounds [9]HBNG and [11]HBNG. In both cases, the resulting crystals belonged to the centrosymmetric triclinic P-1 space group, and consequently contained both isomers. For both nanographenes, the packing displays voids between the HBNGs that contain highly disordered solvent molecules; some of them could be modelled as oxygen atoms from water molecules in [9]HBNG, and dichloroethane and oxygen atoms from water in [11]HBNG. The crystals were found to be extremely unstable out of their mother liquors due to the loss of these interstitial solvent molecules. Additionally, many of the tBu groups are also heavily disordered and were modelled with two alternative sets of positions. Table AP3.1. Sample and crystal data for [9]HBNG. CCDC code 2233478 Chemical formula (C134H118)· 4.38 H2O Formula weight 1807.30 g/mol Temperature 250(2) K Wavelength 0.71073 Å Crystal size 0.026 x 0.236 x 0.507 mm Crystal habit dark brown-orange plate Crystal system triclinic Space group P -1 Unit cell dimensions a = 13.382(4) Å α = 98.887(11)° b = 17.848(5) Å β = 92.864(13)° c = 28.398(8) Å γ = 109.273(11)° Volume 6288.(3) Å3 Z 2 Density (calculated) 0.955 g/cm3 Absorption coefficient 0.056 mm-1 F(000) 1914 Table AP3.2. Data collection and structure refinement for [9]HBNG. Theta range for data collection 1.54 to 25.36° Index ranges -15<=h<=15, -20<=k<=21, -34<=l<=34 Reflections collected 217729 Independent reflections 22608 [R(int) = 0.0631] Coverage of independent reflections 98.0% Absorption correction Multi-Scan Appendix 3. Chapter 3 182 Max. and min. transmission 1.0000 and 0.9300 Refinement method Full-matrix least-squares on F2 Refinement program SHELXL-2019/1 (Sheldrick, 2019) Function minimized Σ w(Fo 2 - Fc 2)2 Data / restraints / parameters 22608 / 1461 / 1386 Goodness-of-fit on F2 1.028 Final R indices 13523 data; I>2σ(I) R1 = 0.1277, wR2 = 0.3571 all data R1 = 0.1900, wR2 = 0.4151 Weighting scheme w=1/[σ2(Fo 2)+(0.2000P)2+22.7000P]; where P=(Fo 2+2Fc 2)/3 Largest diff. peak and hole 1.687 and -0.662 eÅ-3 R.M.S. deviation from mean 0.251 eÅ-3 Table AP3.3. Sample and crystal data for [11]HBNG. CCDC code 2233479 Chemical formula (C142H122)·1.5(C2H4Cl2)·5 H2O Formula weight 2091.02 g/mol Temperature 150(2) K Wavelength 0.71073 Å Crystal size 0.189 x 0.223 x 0.235 mm Crystal habit dark brown prism Crystal system triclinic Space group P -1 Unit cell dimensions a = 18.7384(8) Å α = 111.0735(15)° b = 20.1754(8) Å β = 108.7063(15)° c = 20.9971(8) Å γ = 102.5524(16)° Volume 6500.0(5) Å3 Z 2 Density (calculated) 1.069 g/cm3 Absorption coefficient 0.126 mm-1 F(000) 2182 Table AP3.4. Data collection and structure refinement for [11]HBNG. Theta range for data collection 1.68 to 25.35° Index ranges -22<=h<=22, -24<=k<=24, -25<=l<=25 Reflections collected 101650 Independent reflections 23790 [R(int) = 0.0383] Coverage of independent reflections 99.9% Absorption correction Multi-Scan Max. and min. transmission 0.9760 and 0.9710 Appendix 3. Chapter 3 183 Structure solution technique direct methods Structure solution program XT, VERSION 2018/2 Refinement method Full-matrix least-squares on F2 Refinement program SHELXL-2019/1 (Sheldrick, 2019) Function minimized Σ w(Fo 2 - Fc 2)2 Data / restraints / parameters 23790 / 1542 / 1538 Goodness-of-fit on F2 1.052 Final R indices 15860 data; I>2σ(I) R1 = 0.1182, wR2 = 0.3678 all data R1 = 0.1630, wR2 = 0.4228 Weighting scheme w=1/[σ2(Fo 2)+(0.3000P)2+7.9001P]; where P=(Fo 2+2Fc 2)/3 Largest diff. peak and hole 2.195 and -0.667 eÅ-3 R.M.S. deviation from mean 0.220 eÅ-3 Figure AP3.1. Centroids in the overlapping rings in molecules [9]HBNG, [10]HBNG and [11]HBNG. Top, zenithal view; bottom, lateral view. Appendix 3. Chapter 3 184 Photophysical Details Table AP3.5. Data from the (multi)exponential fittings of the fluorescence decays of helical bilayer nanographenes [9]HBNG, [10]HBNG and [11]HBNG, obtained at different emission detection wavelengths (peaks or shoulders, sh) with excitation at 405 nm. Comp. detect (nm) τ1 (ns) B1 I1 τ2 (ns) B2 I2 τ3 (ns) B3 I3 τInt (ns) τAmp (ns) [9]HBNG 512 (sh) 11.2 0.51 0.79 3.2 0.49 0.21 - - - 9.5 7.3 em max = 575 nm 578 (peak) 14.3 0.71 0.91 3.2 0.29 0.09 - - - 13.4 11.1 620 (sh) 14.3 0.66 0.90 3.1 0.34 0.10 - - - 13.2 10.5 [10]HBNG 507 (sh) 14.6 0.31 0.78 5.1 0.20 0.18 0.52 0.48 0.04 12.3 5.8 em max = 543 nm 540 (peak) 13.2 0.40 0.90 3.4 0.15 0.09 0.11 0.45 0.01 12.2 5.8 570 (sh) 13.0 0.48 0.86 3.9 0.24 0.13 0.29 0.28 0.01 11.7 7.3 630 (sh) 12.3 0.25 0.75 3.7 0.28 0.24 0.12 0.47 0.01 10.1 4.2 [11]HBNG em max = 528 nm 530 (peak) 8.7 1.0 - - - - - - - - - 560 (sh) 8.7 1.0 - - - - - - - - - Table AP3.6. Estimated radiative (kr) and nonradiative (knr) deactivation rate constants of the singlet excited states calculated with data from the (multi)exponential fitting of the fluorescence decays of the helical bilayer nanographenes [9]HBNG, [10]HBNG and [11]HBNG, and their respective emission quantum yields. Compound Amp (ns) em kr S1→S0 (s-1)a knr S1→S0 (s-1)b [9]HBNG 11.1 0.22 2.0  107 7.0  107 [10]HBNG 5.8 0.10 1.7  107 1.5  108 [11]HBNG 8.7 0.11 1.3  107 1.0  108 akr S1→S0 =  em/em. b knr S1→S0 = (1/em) – kr S1→S0 Appendix 3. Chapter 3 185 Chiral HPLC of HBNGs Racemic bilayer helicenes ([9]HBNG, [10]HBNG, [11]HBNG) were resolved by semipreparative CSP HPLC using Chiralpak IE column with a mixture of heptane and 1% isopropyl alcohol in toluene as mobile phase. Unfortunately, a high loss of material was observed during the chromatographic separations, which may indicate irreversible adhesion on the column. Below, the experimental conditions of the HPLC separations and characterization of the obtained enantioenriched samples are summarized. Analytical separations were performed on Chiralpak IE (250 x 4.6 mm, 5 µm, DAICEL) column using isocratic elution with heptane –5% THF in toluene 60:40 at 1 mL/min, semi-preparative separations were done using a Chiralpak IE (250 x 20 mm, 5 µm, DAICEL) column and a gradient elution with heptane– 1% isopropyl alcohol in toluene [80:20] to [50:50] at 20 mL/min. Injections of cca 10 mg / 1.5 mL of the racemic material in a mixture of 1,1,2,2-tetrachloroethane and heptane (1:2) per run were performed. Figure AP3.2. HPLC chromatogram corresponding to the isolated enantiomers of [9]HBNG. Appendix 3. Chapter 3 186 Figure AP3.3. HPLC chromatogram corresponding to the isolated enantiomers of [10]HBNG. Appendix 3. Chapter 3 187 Figure AP3.4. HPLC chromatogram corresponding to the isolated enantiomers of [11]HBNG. Table AP3.7. Optical rotation values of the enantioenriched samples. Compound Optical rotation a [α]D20 HPLC enantiomeric purity e.e. (%) NMR purityb (wt%) (+)-[9]HBNG +6706 98 96 (−)-[9]HBNG −7120 99 97 (+)-[10]HBNG +2447 73 94 (−)-[10]HBNG −4480 99 93 (+)-[11]HBNG +4033 99 92 (−)-[11]HBNG −4270 99 92 a Optical rotations were measured in THF solutions (c = 0.078 – 0.095 g/100mL) b Despite drying in vacuo, residual solvents (mainly toluene and/or dichloromethane) were found to be present in the resolved samples by 1H NMR (in THF-d8). Appendix 3. Chapter 3 188 Vibrational Raman Spectroscopy Figure AP3.5. Experimental (black solid line; λexc = 325 nm) and theoretical (grey solid line) Raman spectra of [9]HBNG. Theoretical Raman spectrum was calculated at the optimized CAM-B3LYP/3-21G* geometry and scaled down uniformly by a factor of 0.96. G Band 1585 cm-1 D Band 1305 cm-1 G Band 1609 cm-1 Appendix 3. Chapter 3 189 Figure AP3.6. Vibrational normal coordinates corresponding to G and D bands in the computed Raman spectrum of [9]HBNG (Figure S19). Bottom graphene layer is showed in light grey for clarity. G Band 1585 cm-1 D Band 1305 cm-1 G Band 1609 cm-1 Appendix 4 Appendix 4. Chapter 4 193 APPENDIX 4. Chapter 4 Crystallographic Data Single clear colorless prism-shaped crystals of 26b were obtained from a methanol- dichloromethane mixture. A suitable crystal was selected and placed on a MiTeGen micromount on an XtaLAB Synergy R, HyPix-Arc 100 diffractometer. The structure was solved with the ShelXT (Sheldrick, 2015) structure solution program using the Intrinsic Phasing solution method and by using Olex2 (Dolomanov et al., 2009) as the graphical interface. The model was refined with version 2018/3 of ShelXL 2018/3 (Sheldrick, 2015) using Least Squares minimization. Table AP4.1. Data collection and structure refinement for 26b. Formula C141.44 H151.31Cl4O 5.75S Dcalc/ g·cm−3 1.029 μ/mm−1 1.302 Formula Weight 2117.08 Color Clear colorless Shape Prism Size/mm3 0.26×0.14×0.08 T/K 200.00(10) Crystal System Orthorhombic Flack Parameter 0.081(7) Hooft Parameter 0.098(5) Space Group P212121 a/Å 18.2270(2) b/Å 21.8090(3) c/Å 34.3878(5) α/° 90 β/° 90 γ/° 90 V/Å3 13669.6(3) Z 4 Z’ 1 Wavelength/Å 1.54184 Radiation type Cu Kα Θmin/° 2.399 Θmax/° 68.528 Measured Refl. 125896 Independent Refl. 25016 Reflections with I > 2(I) 20392 Rint 0.0758 Parameters 1409 Restraints 1328 Appendix 4. Chapter 4 194 Largest Peak 0.919 Deepest Hole −0.551 GooF 1.058 wR2 (all data) 0.4032 wR2 0.3855 R1 (all data) 0.1514 R1 0.1414 Single clear intense orange plate-shaped crystals of P-oxa[9]HBNG were obtained from methanol. A suitable crystal 0.17×0.14×0.05 mm3 was selected and placed on a MiTeGen micromount on an XtaLAB Synergy R, HyPix-Arc 100 diffractometer. The structure was solved with the ShelXT 2018/2 (Sheldrick, 2018) structure solution program using the Intrinsic Phasing solution method and by using Olex2 (Dolomanov et al., 2009) as the graphical interface. The model was refined with version 2018/3 of ShelXL 2018/3 (Sheldrick, 2015) using Least Squares minimization. Table AP4.2. Data collection and structure refinement for P-oxa[9]HBNG. Formula C132 H116O8 Dcalc/ g·cm−3 0.922 μ/mm−1 0.436 Formula Weight 1830.24 Color Clear intense orange Shape Plate Size/mm3 0.17×0.14×0.05 T/K 200.00(10) Crystal System Orthorhombic Flack Parameter −2.6(2) Hooft Parameter −2.97(16) Space Group C2221 a/Å 17.1383(3) b/Å 30.5376(4) c/Å 25.1957(7) α/° 90 β/° 90 γ/° 90 V/Å3 13186.5(5) Z 4 Z’ 0.5 Wavelength/Å 1.54184 Radiation type Cu Kα Θmin/° 2.894 Θmax/° 68.610 Measured Refl. 61631 Appendix 4. Chapter 4 195 Independent Refl. 12073 Reflections with I > 2(I) 8805 Rint 0.0577 Parameters 662 Restraints 37 Largest Peak 1.170 Deepest Hole −0.707 GooF 1.041 wR2 (all data) 0.4395 wR2 0.4138 R1 (all data) 0.1882 R1 0.1701 Single intense brown prism-shaped crystals of M-oxa[9]HBNG were obtained from methanol. A suitable crystal 0.27×0.14×0.07 mm3 was selected and placed on a MiTeGen micromount on an XtaLAB Synergy R, HyPix-Arc 100 diffractometer. The structure was solved with the ShelXT 2018/2 (Sheldrick, 2018) structure solution program using the Intrinsic Phasing solution method and by using Olex2 (Dolomanov et al., 2009) as the graphical interface. The model was refined with version 2018/3 of ShelXL 2018/3 (Sheldrick, 2015) using Least Squares minimization. Table AP4.3. Data collection and structure refinement for M-oxa[9]HBNG. Formula C132H116O2.25 Dcalc/ g·cm−3 0.849 μ/mm−1 0.371 Formula Weight 1738.24 Color Intense brown Shape Prism Size/mm3 0.27×0.14×0.07 T/K 200.00(10) Crystal System Orthorhombic Flack Parameter −1.31(18) Hooft Parameter −1.59(12) Space Group C2221 a/Å 17.1026(2) b/Å 30.4016(3) c/Å 26.1587(4) α/° 90 β/° 90 γ/° 90 V/Å3 13601.1(3) Z 4 Z’ 0.5 Appendix 4. Chapter 4 196 Wavelength/Å 1.54184 Radiation type Cu Kα Θmin/° 2.907 Θmax/° 68.358 Measured Refl. 63965 Independent Refl. 12453 Reflections with I > 2(I) 9674 Rint 0.0270 Parameters 690 Restraints 779 Largest Peak 1.320 Deepest Hole −0.546 GooF 1.029 wR2 (all data) 0.3144 wR2 0.2975 R1 (all data) 0.1335 R1 0.1208 Chiral HPLC of oxa[9]HBNG enantiomers The enantiomers were resolved by chiral HPLC using a column Chiralpack IC 5μm, 98% hexane 2% toluene:isopropanol 4:1, recording the absorption by UV-vis and CD at 370 nm. The negative cotton effect for M-oxa[9]HBNG reveals an e.e.>99%. The cotton effects in the CD chromatogram of P-oxa[9]HBNG shows an e.e.=99%. Appendix 4. Chapter 4 197 Figure AP4.1. HPLC chromatogram of enantiopure (P)- and (M)-oxa[9]HBNG recorded by UV-vis and CD at 370 nm. 0 1 2 3 4 5 6 7 8 9 10 11 U V O p ti c a l d e n s it y Time (min) (P)-oxa[9]HBNG (M)-oxa[9]HBNG Appendix 5 Appendix 5. Spectra 201 APPENDIX 5. Spectra Chapter 1 Figure S1. 1H NMR spectra for compound 4. 1H-NMR, 700 MHz, Chloroform-d 4 Appendix 5. Spectra 202 1H-NMR, 300 MHz, Chloroform-d 13C-NMR, 75 MHz, Chloroform-d DEPT, 75 MHz, Chloroform-d 6a Appendix 5. Spectra 203 1H-NMR, 700 MHz, Methylene Chloride-d2 13C-NMR, 176 MHz , Methylene Chloride-d2 7 Appendix 5. Spectra 204 1H-NMR, 300 MHz, Benzene-d6 8 Appendix 5. Spectra 205 9 1H-NMR, 500 MHz, Chloroform-d Appendix 5. Spectra 206 9 1H-NMR, 500 MHz, Chloroform-d Appendix 5. Spectra 207 1H-NMR, 500 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d DEPT, 126 MHz, Chloroform-d 10 Appendix 5. Spectra 208 1H-NMR, 300 MHz, Benzene-d6 13C-NMR{1H}{19F}, 126 MHz, Benzene-d6 19F-NMR, 282 MHz, Benzene-d6 12 Appendix 5. Spectra 209 1H-NMR, 500 MHz, Chloroform-d 13C-NMR{1H}{19F}, 126 MHz, Chloroform-d 19F-NMR, 471 MHz, Chloroform-d 6b Appendix 5. Spectra 210 1H-NMR, 500 MHz, Chloroform-d 13C-NMR{1H}{19F}, 126 MHz, Chloroform-d 19F-NMR, 471 MHz, Chloroform-d 13 Appendix 5. Spectra 211 Chapter 2 1H-NMR, 300 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d 13C-NMR{1H}{19F}, 126 MHz, Chloroform-d 15a Appendix 5. Spectra 212 19F-NMR, 471 MHz, Chloroform-d 15a DEPT, 126 MHz, Chloroform-d Appendix 5. Spectra 213 1H-NMR, 300 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d 13C-NMR{1H}{19F}, 126 MHz, Chloroform-d 15b 19F-NMR, 471 MHz, Chloroform-d Appendix 5. Spectra 214 1H-NMR, 300 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d 13C-NMR{1H}{19F}, 126 MHz, Chloroform-d 15c Appendix 5. Spectra 215 19F-NMR, 471 MHz, Chloroform-d 15c Appendix 5. Spectra 216 1H-NMR, 500 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d 13C-NMR{1H}{19F}, 126 MHz, Chloroform-d 16a Appendix 5. Spectra 217 19F-NMR, 471 MHz, Chloroform-d 16a Appendix 5. Spectra 218 1H-NMR, 500 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d 16b Appendix 5. Spectra 219 13C-NMR{1H}{19F}, 126 MHz, Chloroform-d 19F-NMR, 282 MHz, Chloroform-d 16b Appendix 5. Spectra 220 1H-NMR, 500 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d 16c Appendix 5. Spectra 221 13C-NMR{1H}{19F}, 126 MHz, Chloroform-d 19F-NMR, 471 MHz, Chloroform-d 16c Appendix 5. Spectra 222 1H-NMR, 300 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d 17a Appendix 5. Spectra 223 13C-NMR{1H}{19F}, 126 MHz, Chloroform-d 19F-NMR, 471 MHz, Chloroform-d 17a Appendix 5. Spectra 224 1H-NMR, 500 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d 17b Appendix 5. Spectra 225 13C-NMR{1H}{19F}, 126 MHz, Chloroform-d 19F-NMR, 471 MHz, Chloroform-d 17b Appendix 5. Spectra 226 1H-NMR, 500 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d 17c Appendix 5. Spectra 227 13C-NMR{1H}{19F}, 126 MHz, Chloroform-d 19F-NMR, 471 MHz, Chloroform-d 17c Appendix 5. Spectra 228 1H-NMR, 500 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d 19F-NMR, 471 MHz, Chloroform-d 17c’ Appendix 5. Spectra 229 17c’. H,H-COSY 17c’. H,H-COSY Appendix 5. Spectra 230 17c’. H,H-NOESY 17c’. H,H-NOESY Appendix 5. Spectra 231 17c’. H,H-NOESY 17c’. H,H-NOESY Appendix 5. Spectra 232 17c’ Appendix 5. Spectra 233 1H NMR of syn/anti isomeric mixture of 17c COSY between 8.23 (d, J = 8.8 Hz) and 7.61 (dd, J = 8.8, 1.8 Hz) of isomer A (blue), the multiplicities of these signals only are posible in the aromatic ring that contains a tBu group with unsubstituted meta position 17c Appendix 5. Spectra 234 Furthemore, also exists COSY between 8.15 (d, J = 8.8 Hz) and 7.54 (m) of isomer B 17c 17c Appendix 5. Spectra 235 The tBu groups of both isomers also have nOe with protons at 8.81 (d, J = 1.8 Hz) and 8.80 (d, J = 1.8 Hz), one for each isomer, but it is very difficult to determine the crossing-points of these signals Additionally, four protons, 8.81 (d, J = 1.8 Hz), 8.80 (d, J = 1.8 Hz), 7.61 (dd, J = 8.8, 1.8 Hz) and 7. 54 (dd, J = (.8, 1.8 Hz) have HMBC with the same cuaternary carbon (35.1) of a tBu group (two carbons, both isomers) 17c 17c Appendix 5. Spectra 236 Two spin systems: tBu substituted phenyl group (pink, protons assigned for each isomer); unsubtituted phenyl group (orange, unassigned) 17c 17c Appendix 5. Spectra 237 nOe between 8.32 in the unsubtituted phenyl ring and 8.15 in tBu substituted phenyl ring. This nOe only can be possible with proton 12 of anti isomer 17c Appendix 5. Spectra 238 Chapter 3 1H-NMR, 400 MHz, Chloroform-d 13C-NMR, 101 MHz, Chloroform-d 19a Appendix 5. Spectra 239 1H-NMR, 400 MHz, Chloroform-d 13C-NMR, 101 MHz, Chloroform-d 19b Appendix 5. Spectra 240 1H-NMR, 300 MHz, Chloroform-d 13C-NMR, 75 MHz, Chloroform-d DEPT, 75 MHz, Chloroform-d 20a Appendix 5. Spectra 241 1H-NMR, 300 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d DEPT, 126 MHz, Chloroform-d 20b Appendix 5. Spectra 242 1H-NMR, 300 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d DEPT, 126 MHz, Chloroform-d 21a Appendix 5. Spectra 243 1H-NMR, 500 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d DEPT, 126 MHz, Chloroform-d 21b Appendix 5. Spectra 244 1H-NMR, 500 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d DEPT, 126 MHz, Chloroform-d [9]HBNG Appendix 5. Spectra 245 1H-NMR, 700 MHz, Chloroform-d 13C-NMR, 176 MHz, Chloroform-d DEPT, 176 MHz, Chloroform-d [11]HBNG Appendix 5. Spectra 246 Chapter 4 1H-NMR, 300 MHz, Chloroform-d 23 Appendix 5. Spectra 247 1H-NMR, 500 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d DEPT, 126 MHz, Chloroform-d 24 Appendix 5. Spectra 248 1H-NMR, 500 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d DEPT, 126 MHz, Chloroform-d 25 Appendix 5. Spectra 249 26a 1H-NMR, 500 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d DEPT, 126 MHz, Chloroform-d Appendix 5. Spectra 250 26b 1H-NMR, 500 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d DEPT, 126 MHz, Chloroform-d Appendix 5. Spectra 251 0 .8 4 0 .5 8 0 .8 1 0 .6 0 7 .0 2 7 .0 7 1H-NMR, 500 MHz, Chloroform-d Low field signals of compounds 26a (blue) and 26b (orange) High field signals of compounds 26a (blue) and 26b (orange) Appendix 5. Spectra 252 1H-NMR, 500 MHz, Chloroform-d 13C-NMR, 126 MHz, Chloroform-d DEPT, 126 MHz, Chloroform-d Oxa[9]HBNG References References 257 REFERENCES 1. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva, A. A. Fisov, Electric Field Effect in Atomically Thin Carbon Films, Science 2004, 306, 666. 2. P. R. Wallace, The Band Theory of Graphite, Phys. Rev. 1947, 71, 622. 3. J. C. Slonczewski, Band Structure of Graphite, Phys. Rev. 1958, 109, 272. 4. G. W. Semenoff, Condensed-Matter Simulation of a Three-Dimensional Anomaly, Phys. Rev. Lett. 1984, 53, 2449. 5. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, M. I. Katsnelson, I. V. Grigorieva, S. V. Dubonos, A. A. Firsov, Two-dimensional gas of massless Dirac fermions in graphene, Nature 2005, 438, 197. 6. A. Geim, K. Novoselov, The Rise of graphene, Nat. Mater. 2007, 6, 183. 7. A. S. Mayorov, Micrometer-Scale Ballistic Transport in Encapsulated Graphene at Room Temperature, Nano Lett. 2011, 11, 2396. 8. S. V. Morozov, Giant Intrinsic Carrier Mobilities in Graphene and Its Bilayer, Phys. Rev. Lett. 2008, 100, 016602. 9. C. Lee, X. D. Wei, J. W. Kysar, J. Hone, Measurement of the Elastic Properties and Intrinsic Strength of Monolayer Graphene, Science 2008, 321, 385. 10. A. A. Balandin, Thermal Properties of Graphene and Nanostructured Carbon Materials, Nat. Mater. 2011, 10, 569. 11. J. S. Bunch, S. S. Verbridge, J. S. Alden, A. M. van der Zande, J. M. Parpia, H. G. Craighead, P. L. McEuen, Impermeable Atomic Membranes from Graphene Sheets, Nano Lett. 2008, 8, 2458. 12. J. Moser, A. Barreiro, A. Bachtold, Current-Induced Cleaning of Graphene, Appl. Phys. Lett. 2007, 91, 163513. 13. A. K. Worku, D. W. Ayele, Recent Advances of Graphene-Based Materials for Emerging Technologies, Results in Chemistry 2023, 5, 100971. 14. A. Gutiérrez-Cruz, A. R. Ruiz-Hernández, J. F. Vega-Clemente, D. G. Luna-Gazcón, J. Campos-Delgado, A Review of Top-Down and Bottom-Up Synthesis Methods for the Production of Graphene, Graphene Oxide and Reduced Graphene Oxide, J. Mater. Sci. 2022, 57, 14543. 15. K. S. Novoselov, V. I. Fal’ko, L. Colombo, P. R. Gellert, M. G. Schwab, K. Kim, A Roadmap for Graphene, Nature 2012, 490, 192. 16. K. Parvez, S. Yang, X. Feng, K. Müllen, Exfoliation of Graphene via Wet Chemical Routes, Synth. Met. 2015, 210, 123. References 258 17. N. Liu, F. Luo, H. Wu, Y. Liu, C. Zhang, J. Chen, One-Step Ionic-Liquid-Assisted Electrochemical Synthesis of Ionic-Liquid-Functionalized Graphene Sheets Directly from Graphite, Adv. Funct. Mater. 2008, 18, 1518. 18. K. S. Kim, Y. Zhao, H. Jang, S. Y. Lee, J. M. Kim, K. S. Kim, J.-H. Ahn, P. Kim, J.-Y. Choi, B. H. Hong, Large-Scale Pattern Growth of Graphene Films for Stretchable Transparent Electrodes, Nature 2009, 457, 706. 19. X. Li, W. C. J. An, S. K. J. Nah, D. Yang, R. Piner, A. Velamakanni, I. Jung, E. Tutuc, S. K. Banerjee, L. Colombo, R. S. Ruoff, Large-Area Synthesis of High-Quality and Uniform Graphene Films on Copper Foils, Science 2009, 324, 1312. 20. C. Berger, Z. Song, T. Li, X. Li, A. Y. Ogbazghi, R. Feng, Z. Dai, A. N. Marchenkov, E. H. Conrad, P. N. First, W. A. De Heer, Ultrathin Epitaxial Graphite:  2D Electron Gas Properties and a Route toward Graphene-based Nanoelectronics, J. Phys. Chem. B 2004, 108, 19912. 21. J. Lin, Z. Peng, Y. Liu, F. Ruiz-Zepeda, R. Ye, E. L. G. Samuel, M. J. Yacaman, B. I. Yakobson, J. M. Tour, Laser-Induced Porous Graphene Films from Commercial Polymers, Nat. Commun. 2014, 5, 5714. 22. M. Qian, Y. S. Zhou, Y. Gao, J. B. Park, T. Feng, S. M. Huang, Z. Sun, L. Jiang, Y. F. Lu, Formation of Graphene Sheets Through Laser Exfoliation of Highly Ordered Pyrolytic Graphite, Appl. Phys. Lett. 2011, 98, 173108. 23. D. X. Luong, K. V. Bets, W. A. Algozeeb, M. G. Stanford, C. Kittrell, W. Chen, R. V. Salvatierra, M. Ren, E. A. McHugh, P. A. Advincula, Z. Wang, M. Bhatt, H. Guo, V. Mancevski, R. Shahsavari, B. I. Yakobson, J. M. Tour, Gram-Scale Bottom-up Flash Graphene Synthesis, Nature 2020, 577, 647. 24. A. Adetayo, D. Runsewe, Synthesis and Fabrication of Graphene and Graphene Oxide: A Review, Open J. Compos. Mater. 2019, 9, 207. 25. Y. Zhu, S. Murali, W. Cai, X. Li, J. W. Suk, J.R. Potts, R.S. Ruoff, Graphene and Graphene Oxide: Synthesis, Properties, and Applications, Adv. Mater. 2010, 22, 3906. 26. N. Panwar, A. M. Soehartono, K. K. Chan, S. Zeng, G. Xu, J. Qu, P. Coquet, K.-T. Yong, X. Chen, Nanocarbons for Biology and Medicine: Sensing, Imaging, and Drug Delivery, Chem. Rev. 2019, 119, 9559. 27. L. Lin, H. Peng, Z. Liu, Synthesis Challenges for Graphene Industry, Nat. Mat. 2019, 18, 520. 28. A. N. Grigorenko, M. Polini, K. S. Novoselov, Graphene Plasmonics, Nat. Photonics 2012, 6, 749. 29. J. Wu, H. Lin, D. J. Moss, K. P. Loh, B. Jia, Graphene Oxide for Photonics, Electronics and Optoelectronics, Nat. Rev. Chem. 2023, 7, 162. 30. F. Bonaccorso, Z. Sun, T. Hasan, A. C. Ferrari, Graphene Photonics and Optoelectronics, Nat. Photonics 2010, 4, 611. References 259 31. H. Su, Y. H. Hu, Recent Advances in Graphene-Based Materials for Fuel Cell Applications, Energy Sci. Eng. 2021, 9, 958. 32. F. Bonaccorso, L. Colombo, G. Yu,M. Stoller, V. Tozzini, A. C. Ferrari, R. S. Ruoff, V. Pellegrini, Graphene, Related Two-Dimensional Crystals, and Hybrid Systems for Energy Conversion and Storage, Science 2015, 347, 1246501. 33. M. Huang, Z. Li, H. Zhu, Recent Advances of Graphene and Related Materials in Artificial Intelligence, Adv. Intell. Syst. 2022, 4, 2200077. 34. J. Li, H. Zeng, Z. Zeng, Y. Zeng, T. Xie, Promising Graphene-Based Nanomaterials and Their Biomedical Applications and Potential Risks: A Comprehensive Review, ACS Biomater. Sci. Eng. 2021, 7, 5363. 35. E. Galvagno, E. Tartaglia, M. Stratigaki, C. Tossi, L. Marasco, F. Menegazzo, C. Zanardi, F. Omenetto, C. Coletti, A. Traviglia, M. Moglianetti, Present Status and Perspectives of Graphene and Graphene-related Materials in Cultural Heritage, Adv. Funct. Mater. 2024, 2313043. 36. M. T. U. Safian, K. Umar, M. N. Mohamad Ibrahim, Synthesis and Scalability of Graphene and its Derivatives: A Journey Towards Sustainable and Commercial Material, J. Clean. Prod. 2021, 318, 128603. 37. G. Li, A. Luican, J. M. B. Lopes dos Santos, A. H. Castro Neto, A. Reina, J. Kong, E. Y. Andrei, Observation of Van Hove Singularities in Twisted Graphene Layers, Nat. Phys. 2010, 6, 109. 38. L. Van Hove, The Occurrence of Singularities in the Elastic Frequency Distribution of a Crystal, Phys. Rev. 1953, 89, 1189. 39. Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi, E. Kaxiras, P. Jarillo-Herrero, Unconventional Superconductivity in Magic-angle Graphene Superlattices, Nature 2018, 556, 43. 40. E. Y. Andrei, D. K. Efetov, P. Jarillo- Herrero, A. H. MacDonald, K. Fai Mak, T. Senthil, E. Tutuc, A. Yazdani, A. F. Young, The Marvels of Moiré Materials, Nat. Rev. Mater. 2021, 6, 201. 41. A. Halleux, R. H. Martin, G. S. D. King,Helv. Synthèses dans la série des dérivés polycycliques aromatiques hautement condensés. L'hexabenzo-1,12; 2,3; 4,5; 6,7; 8,9; 10,11-coronène, le tétrabenzo-4,5; 6,7; 11,12; 13,14-péropyrène et le tétrabenzo-1,2; 3,4; 8,9; 10,11-bisanthène, Chim. Acta 1958, 41, 1177. 42. L. Chen, Y. Hernandez, X. Feng, K. Müllen, From Nanographene and Graphene Nanoribbons to Graphene Sheets: Chemical Synthesis, Angew. Chem. Int. Ed. 2012, 51, 7640. 43. J. Liu, X. Feng, Synthetic Tailoring of Graphene Nanostructures with Zigzag-Edged Topologies: Progress and Perspectives, Angew. Chem. Int. Ed. 2020, 59, 23386. 44. C. D. Simpson, J. D. Brand, A. J. Berresheim, L. Przybilla, H. J. Räder, K. Müllen, Synthesis of a Giant 222 Carbon Graphite Sheet, Chem. Eur. J. 2002, 8, 1424. References 260 45. Z. Liu, S. Fu, X. Liu, A. Narita, P. Samori, M. Bonn, H. I. Wang, Small Size, Big Impact: Recent Progress in Bottom-Up Synthesized Nanographenes for Optoelectronic and Energy Applications, Adv. Sci. 2022, 9, 2106055. 46. Y. Gu, Z. Qiu, K. Müllen, Nanographenes and Graphene Nanoribbons as Multitalents of Present and Future Materials Science, J. Am. Chem. Soc. 2022, 144, 11499. 47. A. Narita, X.-Y. Wang, X. Feng, K. Müllen, New Advances in Nanographene Chemistry, Chem. Soc. Rev. 2015, 44, 6616. 48. S. Song, J. Su, M. Telychko, J. Li, G. Li, Y. Li, C. Su, J. Wu, J. Lu, On-surface Synthesis of Graphene Nanostructures with π-Magnetism, Chem. Soc. Rev. 2021, 50, 3238. 49. Y. Segawa, H. Ito, K. Itami, Structurally Uniform and Atomically Precise Carbon Nanostructures, Nat. Rev. Mater. 2016, 1, 15002. 50. (a) R. Scholl, C. Seer, R. Weitzenbök, Perylen, ein hoch kondensierter aromatischer Kohlenwasserstoff C20H12, Chem. Ber. 1910, 43, 2202. (b) R. Scholl, C. Seer, Abspaltung aromatisch gebundenen Wasserstoffs und Verknüpfung aromatischer Kerne durch Aluminiumchlorid, Liebigs Ann. Chem. 1912, 394, 111. (c) R. Scholl, C. Seer, Abspaltung aromatisch gebundenen Wasserstoffes unter Verknüpfung aromatischer Kerne durch Aluminiumchlorid, 6. Mitteilung: Versuche mit Phenol-äthern und mit Diphenyl-methan, Chem. Ber. 1922, 55, 330. (d) E. Clar, D. G. Stewart, Aromatic Hydrocarbons. LXV. Triangulene Derivatives, J. Am. Chem. Soc. 1953, 75, 2667. (e) E. Clar, W. Schmidt, Localised vs delocalised molecular orbitals in aromatic hydrocarbons, Tetrahedron 1979, 35, 2673. 51. S. Ito, M. Wehmeier, J. D. Brand, C. Kübel, R. Epsch, J. P. Rabe, K. Müllen, Synthesis and Self-Assembly of Functionalized Hexa-peri-hexabenzocoronenes, Chem. Eur. J. 2000, 6, 4327. 52. S. Ma, J. Gu, C. Lin, Z. Luo, Y. Zhu, J. Wang, Supertwistacene: A Helical Graphene Nanoribbon, J. Am. Chem. Soc. 2020, 142, 16887. 53. A. Narita, I. A. Verzhbitskiy, W. Frederickx, K. S. Mali, S. A. Jensen, M. R. Hansen, M. Bonn, S. De Feyter, C. Casiraghi, X. Feng, K. Müllen, Bottom-Up Synthesis of Liquid- Phase-Processable Graphene Nanoribbons with Near-Infrared Absorption, ACS Nano 2014, 8, 11622. 54. J. Urieta-Mora, M. Krug, W. Alex, J. Perles, I. Fernández, A. Molina-Ontoria, D. M. Guldi, N. Martín, Homo and Hetero Molecular 3D Nanographenes Employing a Cyclooctatetraene Scaffold, J. Am. Chem. Soc. 2020, 142, 4162. 55. C. M. Cruz, I. R. Márquez, S. Castro-Fernández, J. M. Cuerva, E. Maçôas, A. G. Campaña, A Triskelion-Shaped Saddle–Helix Hybrid Nanographene, Angew. Chem. Int. Ed. 2019, 58, 8068. 56. K. Kawasumi, Q. Zhang, Y. Segawa, L. T. Scott, K. Itami, A Grossly Warped Nanographene and the Consequences of Multiple Odd-membered-ring Defects, Nat. Chem. 2013, 5, 739. References 261 57. X.-S. Zhang, Y.-Y. Huang, J. Zhang, W. Meng, Q. Peng, R. Kong, Z. Xiao, J. Liu, M. Huang, Y. Yi, L. Chen, Q. Fan, G. Lin, Z. Liu, G. Zhang, L. Jiang, D. Zhang, Dicyclohepta[ijkl,uvwx]rubicene with Two Pentagons and Two Heptagons as a Stable and Planar Non-benzenoid Nanographene, Angew. Chem. Int. Ed. 2020, 59, 3529. 58. V. Akhmetov, M. Feofanov, O. Papaianina, S. Troyanov, K. Amsharov, Towards Nonalternant Nanographenes through Self-Promoted Intramolecular Indenoannulation Cascade by C−F Bond Activation, Chem. Eur. J. 2019, 25, 11609. 59. M. Daigle, A. Picard‐Lafond, E. Soligo, J.-F. Morin, Regioselective Synthesis of Nanographenes by Photochemical Cyclodehydrochlorination, Angew. Chem. Int. Ed. 2016, 55, 2042. 60. L. T. Scott, E. Jackson, Q. Zhang, B. D. Steinberg, M. Bancu, B. Li, A Short, Rigid, Structurally Pure Carbon Nanotube by Stepwise Chemical Synthesis, J. Am. Chem. Soc. 2012, 134, 107. 61. W.-J. Kong, L. H. Finger, J. C. A. Oliveira, L. Ackermann, Rhodaelectrocatalysis for Annulative C−H Activation: Polycyclic Aromatic Hydrocarbons through Versatile Double Electrocatalysis, Angew. Chem. Int. Ed. 2019, 58, 6342. 62. B. Cirera, A. Sánchez-Grande, B. de la Torre, J. Santos, S. Edalatmanesh, E. Rodríguez- Sánchez, K. Lauwaet, B. Mallada-Faes, R. Zbořil, R. Miranda, O. Gröning, P. Jelínek, N. Martín, D. Écija, Tailoring Topological Order and π-Conjugation to Engineer Quasi- Metallic Polymers, Nat. Nanotech. 2020, 15, 437. 63. (a) R. Scholl, J. Mansfeld, Ber. Dtsch. Chem. Ges. 1910, 43, 1734. (b) R. Scholl, C. Seer, Justus Liebigs, Abspaltung Aromatisch Gebundenen Wasserstoffs und Verknüpfung Aromatischer Kerne durch Aluminiumchlorid, Ann. Chem. 1912, 394, 111. 64. R. S. Jassas, E. U. Mughal, A. Sadiq, R. I. Alsantali, M. M. Al-Rooqi, N. Naeem, Z. Moussa, S. A. Ahmed, Scholl Reaction as a Powerful Tool for the Synthesis of Nanographenes: a Systematic Review, RSC Adv. 2021, 11, 32158. 65. M. Grzybowski, K. Skonieczny, H. Butenschçn, D. T. Gryko, Comparison of Oxidative Aromatic Coupling and the Scholl Reaction, Angew. Chem. Int. Ed. 2013, 52, 9900. 66. M. Grzybowski, B. Sadowski, H. Butenschçn, D. T. Gryko, Synthetic Applications of Oxidative Aromatic Coupling—From Biphenols to Nanographenes, Angew. Chem. Int. Ed. 2020, 59, 2998. 67. G. Baddeley, J. Kenner, The Constitution of Apocinchene and Syntheses of its Methyl and Ethyl Ethers, J. Chem. Soc. 1935, 303. 68. G. Baddeley, Hydrogen Chloride–Aluminium Chloride as an Agent of Isomerization, J. Chem. Soc. 1950, 994. 69. J. Kenner, Biographical Memoirs of Fellows of the Royal Society, J. Soc. Chem. Ind. 1933, 42, 470. References 262 70. L. Zhai, R. Shukla, S. H. Wadumethrige, W. Rathore, Probing the Arenium-Ion (ProtonTransfer) versus the Cation-Radical (Electron Transfer) Mechanism of Scholl Reaction Using DDQ as Oxidant, J. Org. Chem. 2010, 75, 4748. 71. L. Zhai, R. Shukla, R. Rathore, Oxidative C−C Bond Formation (Scholl Reaction) with DDQ as an Efficient and Easily Recyclable Oxidant, Org. Lett. 2009, 11, 3474. 72. N. Ponugoti, V. Parthasarathy, Rearrangements in Scholl Reaction, Chem. Eur. J. 2022, 28, e202103530. 73. Y. Zhang, S. H. Pun, Q. Miao, The Scholl Reaction as a Powerful Tool for Synthesis of Curved Polycyclic Aromatics, Chem. Rev. 2022, 122, 14554. 74. Y. Zou, Y. Han, S. Wu, X. Hou, C. H. E. Chow, J. Wu, Scholl Reaction of Perylene-Based Polyphenylene Precursors under Different Conditions: Formation of Hexagon or Octagon?, Angew. Chem. Int. Ed. 2021, 60, 17654. 75. F. Ko, K. Hirabayashi, T. Shimizu, K.-I. Sugiura, Unexpected Phenyl Group Rearrangement of Thiele’s Hydrocarbon Derivative Under Polycyclic Aromatic Hydrocarbon Synthesis, Tetrahedron 2018, 59, 4251. 76. J. Liu, A. Narita, S. Osella, W. Zhang, D. Schollmeyer, D. Beljonne, X. Feng, K. Müllen, Unexpected Scholl Reaction of 6,7,13,14-Tetraarylbenzo[k]tetraphene: Selective Formation of Five-Membered Rings in Polycyclic Aromatic Hydrocarbons, J. Am. Chem. Soc. 2016, 138, 2602. 77. J. Liu, S. Osella, J. Ma, R. Berger, D. Beljonne, D. Schollmeyer, X. Feng, K. Müllen, Fused Dibenzo[a,m]rubicene: A New Bowl-Shaped Subunit of C70 Containing Two Pentagons, J. Am. Chem. Soc. 2016, 138, 8364. 78. M. Müller, V. S. Iyer, C. Kübel, V. Enkelmann, K. Müllen, Polycyclic Aromatic Hydrocarbons by Cyclodehydrogenation and Skeletal Rearrangement of Oligophenylenes, Angew. Chem. Int. Ed. 1997, 36, 1607. 79. F. E. Golling, M. Quernheim, M. Wagner, T. Nishiuchi, K. Müllen, Concise Synthesis of 3D π-Extended Polyphenylene Cylinders, Angew. Chem. Int. Ed. 2014, 53, 1525. 80. X. Dou, X. Yang, G. J. Bodwell, M. Wagner, V. Enkelmann, K. Müllen, Unexpected Phenyl Group Rearrangement during an Intramolecular Scholl Reaction Leading to an Alkoxy-Substituted Hexa-peri-hexabenzocoronene, Org. Lett. 2007, 9, 2485. 81. M. Müller, V. S. Iyer, C. Kübel, V. Enkelmann, K. Müllen, Polycyclic Aromatic Hydrocarbons by Cyclodehydrogenation and Skeletal Rearrangement of Oligophenylenes, Angew. Chem. Int. Ed. 1997, 36, 1607. 82. A. Pradhan, P. Dechambenoit, H. Bock, F. Durola, Twisted Polycyclic Arenes by Intramolecular Scholl Reactions of C3-Symmetric Precursors, J. Org. Chem. 2013, 78, 2266. 83. Z. Qiu, S. Asako, Y. Hu, C.-W. Ju, T. Liu, L. Rondin, D. Schollmeyer, J.-S. Lauret, K. Müllen, A. Narita, Negatively Curved Nanographene with Heptagonal and [5]Helicene Units, J. Am. Chem. Soc. 2020, 142, 14814. References 263 84. J. Ma, Y. Fu, E. Dmitrieva, F. Liu, H. Komber, F. Hennersdorf, A. A. Popov, J. J. Weigand, J. Liu, X. Feng, Helical Nanographenes Containing an Azulene Unit: Synthesis, Crystal Structures, and Properties, Angew. Chem. Int. Ed. 2020, 59, 5637. 85. J. Ma, Y. Fu, E. Dmitrieva, F. Liu, H. Komber, F. Hennersdorf, A. A. Popov, J. J. Weigand, J. Liu, X. Feng, Helical Nanographenes Containing an Azulene Unit: Synthesis, Crystal Structures, and Properties, Angew. Chem. Int. Ed. 2020, 132, 5686. 86. Y. Han, Z. Xue, G. Li, Y. Gu, Y. Ni, S. Dong, C. Chi, Formation of Azulene-Embedded Nanographene: Naphthalene to Azulene Rearrangement During the Scholl Reaction, Angew. Chem. Int. Ed. 2020, 59, 9026. 87. J. M. Fernández-García, P. J. Evans, S. Medina Rivero, I. Fernández, D. García- Fresnadillo, J. Perles, J. Casado, N. Martín, π-Extended Corannulene-Based Nanographenes: Selective Formation of Negative Curvature, J. Am. Chem. Soc. 2018, 140, 17188. 88. L. T. Bratz and S. von Niementowski, Synthesen des 8.8′-Dioxy-5.5′-dichinolyls und ein Fall der direkten Chlorierung mittels Eisenchlorids, Ber. Dtsch. Chem. Ges. (A and B Ser.), 1919, 52, 189. 89. D. M. Baier, S. Gratz, B. Farhadi Jahromi, S. Hellmann, K. Bergheim, W. Pickhardt, R. Schmid, L. Borchardt, Beyond the Scholl Reaction – One-step Planarization and Edge Chlorination of Nanographenes by Mechanochemistry, RSC Adv., 2021, 11, 38026. 90. J. Stanojkovic, R. William, Z. Zhang, I. Fernández, J. Zhou, R. D. Webster, M. C. Stuparu. Synthesis of Precisely Functionalizable Curved Nanographenes via Graphitization-induced Regioselective Chlorination in a Mechanochemical Scholl Reaction, Nat. Commun. 2023, 14, 803. 91. T. Fujikawa, N. Mitoma, A. Wakamiya, A. Saeki, Y. Segawa, K. Itami, Synthesis, Properties, and Crystal Structures of π-Extended Double [6]helicenes: Contorted Multi- Dimensional Stacking Lattice, Org. Biomol. Chem. 2017, 15, 4697. 92. X. Yang, M. Hoffmann, F. Rominger, T. Kirschbaum, A. Dreuw, M. Mastalerz, Functionalized Contorted Polycyclic Aromatic Hydrocarbons by a One-Step Cyclopentannulation and Regioselective Triflyloxylation, Angew. Chem. Int. Ed. 2019, 58, 10650. 93. X. Yang, F. Rominger, M. Mastalerz, Contorted Polycyclic Aromatic Hydrocarbons with Two Embedded Azulene Units, Angew. Chem. Int. Ed. 2019, 58, 17577. 94. S. Liu, D. Xia, M. Baumgarten, Rigidly Fused Spiro-Conjugated π-Systems, Chem. Plus Chem. 2021, 86, 36. 95. X. Zhang, Z. Xu, W. Si, K. Oniwa, M. Bao, Y. Yamamoto, T. Jin, Synthesis of Extended Polycyclic Aromatic Hydrocarbons by Oxidative Tandem Spirocyclization and 1,2-aryl Migration, Nat. Commun. 2017, 8, 15073. 96. J. L. Ormsby, T. D. Black, C. L. Hilton, Bharat, B. T. King, Rearrangements in the Scholl Oxidation: Implications for Molecular Architectures, Tetrahedron 2008, 64, 11370. References 264 97. S. Mauskopf, Crystals and Compounds: Molecular Structure and Composition in Nineteenth-Century French Science, Trans. Am. Philos. Soc. 1976, 66, 56. 98. S. Mauskopf, Crystals and Compounds: Molecular Structure and Composition in Nineteenth-Century French Science, Trans. Am. Philos. Soc. 1976, 66, 61. 99. J.-B. Biot, The Mathematization of Experimental Physics in Napoleonic France, Mém. Acad. R. Sci. Inst. Fr. 1817, 2, 4. 100. L. Pasteur, C. R. Séances Acad. Sci. 1848, 26, 535. 101. J. H. van’t Hoff, Sur les Formules de Structure dans L'espace, Arch. Neerl. 1874, 9, 1. 102. Lord Kelvin, The Molecular Tactics of a Crystal. The 2nd Robert Boyle Lecture J. Oxford Univ. Junior Sci. Club 1894, 18, 25. 103. (a) R.S. Cahn, C.K. Ingold and V. Prelog. Specification of Configuration About Quadricovalent Asymmetric Atoms J. Chem. Soc. 1951, 612. (b) R.S. Cahn, C.K. Ingold and V. Prelog, The Specification of Asymmetric Configuration in Organic Chemistry, Experientia 1956, 12, 81. 104. (a) R.S. Cahn, C.K. Ingold and V. Prelog. Angew. Chem. Int. Ed.1966, 5, 385. (b) V. Prelog, G. Helmchen, Basic Principles of the CIP-System and Proposals for a Revision, Angew. Chem. Int. Ed. 1982, 21, 567. 105. IUPAC. Compendium of Chemical Terminology, 2nd ed. (the "Gold Book"). Compiled by A. D. McNaught and A. Wilkinson. Blackwell Scientific Publications, Oxford (1997). 106. M. Rickhaus, M. Mayor, M. Juríček, Chirality in Curved Polyaromatic Systems, Chem. Soc. Rev. 2017, 46, 1643. 107. M. Rickhaus, M. Mayor, M. Juríček, Strain-induced Helical Chirality in Polyaromatic Systems, Chem. Soc. Rev. 2016, 45, 1542. 108. K. Oki, M. Takase, S. Mori, A. Shiotari, Y. Sugimoto, K. Ohara, T. Okujima, H. Uno, Synthesis, Structures, and Properties of Core-Expanded Azacoronene Analogue: A Twisted π-System with Two N-Doped Heptagons, J. Am. Chem. Soc. 2018, 140, 10430. 109. J. Liu, B.-W. Li, Y.-Z. Tan, A. Giannakopoulos, C. Sánchez-Sánchez, D. Beljonne, P. Ruffieux, R. Fasel, X. L. Feng, K. Müllen, Toward Cove-Edged Low Band Gap Graphene Nanoribbons, J. Am. Chem. Soc. 2015, 137, 6097. 110. K. Kawasumi, Q. Zhang, Y. Segawa, L. T. Scott, K. Itami, A Grossly Warped Nanographene and the Consequences of Multiple Odd-membered-ring Defects, Nat. Chem. 2013, 5, 739. 111. X. Gu, H. Li, B. Shan, Z. Liu, Q. Miao, Synthesis, Structure, and Properties of Tetrabenzo[7]circulene, Org. Lett. 2017, 19, 2246. 112. S. H. Pun, Y. Wang, M. Chu, C. K. Chan, Y. Li, Z. Liu, Q. Miao, Synthesis, Structures, and Properties of Heptabenzo[7]circulene and Octabenzo[8]circulene, J. Am. Chem. Soc., 2019, 141, 9680. References 265 113. T. Kirschbaum, F. Rominger and M. Mastalerz, A Chiral Polycyclic Aromatic Hydrocarbon Monkey Saddle, Angew. Chem., Int. Ed., 2020, 59, 270. 114. Z. Qiu, S. Asako, Y. Hu, C.-W. Ju, T. Liu, L. Rondin, D. Schollmeyer, J.-S. Lauret, K. Müllen, A. Narita, Negatively Curved Nanographene with Heptagonal and [5]Helicene Units, J. Am. Chem. Soc. 2020, 142, 14814. 115. C. M. Cruz, I. R. Márquez, I. F. A. Mariz, V. Blanco, C. Sánchez-Sánchez, J. M. Sobrado, J. A. Martín-Gago, J. M. Cuerva, E. Maçôas, A. G. Campaña, Enantiopure Distorted Ribbon-shaped Nanographene Combining Two-photon Absorption-based Upconversion and Circularly Polarized Luminescence, Chem. Sci. 2018, 9, 3917. 116. M. A. Medel, R. Tapia, V. Blanco, D. Miguel, S. P. Morcillo, A. G. Campaña, Octagon- Embedded Carbohelicene as a Chiral Motif for Circularly Polarized Luminescence Emission of Saddle-Helix Nanographenes, Angew. Chem. Int. Ed. 2021, 60, 6094. 117. X. Xu, R. Munoz-Marmol, S. Vasylevskyi, A. Villa, G. Folpini, F. Scotognella, G. M. Paterno, A. Narita, Synthesis of Bioctacene-Incorporated Nanographene with Near- Infrared Chiroptical Properties, Angew. Chem. Int. Ed. 2023, 62, e202218350. 118. S. H. Pun, K. M. Cheung, D. Yang, H. Chen, Y. Wang, S. V. Kershaw, Q. Miao, A Near- Infrared Absorbing and Emissive Quadruple Helicene Enabled by the Scholl Reaction of Perylene, Angew. Chem. Int. Ed. 2022, 61, e202113203. 119. M. Krzeszewski, H. Ito, K. Itami, Infinitene: A Helically Twisted Figure-Eight [12]Circulene Topoisomer, Infinitene: A Helically Twisted Figure-Eight [12]Circulene Topoisomer, J. Am. Chem. Soc. 2022, 144, 862. 120. W. Fan, T. Matsuno, Y. Han, X. Wang, Q. Zhou, H. Isobe, J. Wu, Synthesis and Chiral Resolution of Twisted Carbon Nanobelts, J. Am. Chem. Soc. 2021, 143, 15924. 121. W. Fan, T. M. Fukunaga, S. Wu, Y. Han, Q. Zhou, J. Wang, Z. Li, X. Hou, H. Wei, Y. Ni, H. Isobe, J. Wu, Synthesis and Chiral Resolution of a Triply Twisted Möbius Carbon Nanobelt, Nat. Synth. 2023, 2, 880. 122. K. Yin Cheung, S. Gui, C. Deng, H. Liang, Z. Xia, Z. Liu, L. Chi, Q. Miao, Synthesis of Armchair and Chiral Carbon Nanobelts, Chem 2019, 5, 838. 123. Y. Nojima, M. Hasegawa, N. Hara, Y. Imai, Y. Mazaki, Stereogenic Cyclic Oligonaphthalenes Displaying Ring Size-dependent Handedness of Circularly Polarized Luminescence (CPL), Chem. Commun. 2019, 55, 2749. 124. S. Li, R. Li, Y.-K. Zhang, S. Wang, B. Ma, B. Zhang, P. An, BINOL-like Atropisomeric Chiral Nanographene, Chem. Sci. 2023, 14, 3286. 125. S. Mörsel, R. Kellner, A. Hirsch, Thioether-Substituted Hexa-peri-hexabenzocoronenes, Eur. J. Org. Chem. 2023, 26, e202300299. 126. R. Adams, H. C. Yuan, The Stereochemistry of Diphenyls and Analogous Compounds, Chem. Rev. 1933, 12, 261. 127. A. Fechtenkötter, K. Saalwächter, M. A. Harbison, K. Müllen, H. Wolfgang Spiess, Highly Ordered Columnar Structures from Hexa-peri-hexabenzocoronenes—Synthesis, References 266 X-ray Diffraction, and Solid-State Heteronuclear Multiple-Quantum NMR Investigations, Angew. Chem. Int. Ed. 1999, 38, 3039. 128. T. Wöhrle, I. Wurzbach, J. Kirres, A. Kostidou, N. Kapernaum, J. Litterscheidt, J. Christian Haenle, P. Staffeld, A. Baro, F. Giesselmann, S. Laschat, Discotic Liquid Crystals, Chem. Rev. 2016, 116, 1139. 129. J. P. Hill, W. Jin, A. Kosaka, T. Fukushima, H. Ichihara, T. Shimomura, K. Ito, T. Hashizume, N. Ishii, T. Aida, Self-Assembled Hexa-peri-hexabenzocoronene Graphitic Nanotube, Science 2004, 304, 1481. 130. X.-J. Zhao, H. Hou, X.-T. Fan, Y. Wang, Y.-M. Liu, C. Tang, S.-H. Liu, P.-P. Ding, J. Cheng, D.-H. Lin, C. Wang, Y. Yang, Y.-Z. Tan, Molecular Bilayer Graphene, Nat. Commun. 2019, 10, 3057. 131. B. Pigulski, K. Shoyama, M.-J. Sun, F. Würthner, Fluorescence Enhancement by Supramolecular Sequestration of a C54-Nanographene Trisimide by Hexabenzocoronene, J. Am. Chem. Soc. 2022, 144, 5718. 132. X. Yang, M. Brückner, F. Rominger, T. Kirschbaum, M. Mastalerz, Dispersion-driven Formation of Chiral Twisted PAH Double Helices, Chem 2024, 10, 832. 133. A. Oró, F. Romeo-Gella, J. Perles, J. M. Fernández-García, I. Corral, N. Martín, Tetrahedraphene: A Csp3-centered 3D Molecular Nanographene Showing Aggregation- Induced Emission Angew. Chem. Int. Ed. 2023, 62, e202312314. 134. W. Wang, X.-H. Ma, M. Liu, S. Tang, X. Ding, Y. Zhao, Y.-Z. Tan, M. Kertesz, X. Wang, A Triply Negatively Charged Nanographene Bilayer with Spin Frustration, Angew. Chem. Int. Ed. 2023, 135, e202217788. 135. L. Moshniaha, M. Żyła-Karwowska, P. J. Chmielewski, T. Lis, J. Cybińska, E. Gońka, J. Oschwald, T. Drewello, S. Medina Rivero, J. Casado, M. Stępień, Aromatic Nanosandwich Obtained by σ-Dimerization of a Nanographenoid π-Radical, J. Am. Chem. Soc. 2020, 142, 3626. 136. P. J. Evans, J. Ouyang, L. Favereau, J. Crassous, I. Fernández, J. Perles, N. Martín, Synthesis of a Helical Bilayer Nanographene, Angew. Chem. Int. Ed. 2018, 57, 6774. 137. D. Reger, P. Haines, F. W. Heinemann, D. M. Guldi, N. Jux, Oxa[7]superhelicene: A π- Extended Helical Chromophore Based on Hexa-peri-hexabenzocoronenes, Angew. Chem. Int. Ed. 2018, 57, 5938. 138. L. Yang, Y.-Y. Ju, M. A. Medel, Y. Fu, H. Komber, E. Dmitrieva, J.-J. Zhang, S. Obermann, A. G. Campaña, J. Ma, X. Feng, Helical Bilayer Nonbenzenoid Nanographene Bearing a [10]Helicene with Two Embedded Heptagons, Angew. Chem. Int. Ed. 2023, 62, e202216193. 139. Y.-J. Shen, N.-T. Yao, L.-N. Diao, Y. Yang, X.-L. Chen, H.-Y. Gong, A π-Extended Pentadecabenzo[9]Helicene, Angew. Chem. Int. Ed. 2023, 62, e202300840. 140. Y. Zhang, S. Yu, B. Han, Y. Zhou, X. Zhang, X. Gao, Z. Tang, Circularly Polarized Luminescence in Chiral Materials, Matter 2022, 5, 837. References 267 141. Y. Deng, M. Wang, Y. Zhuang, S. Liu, W. Huang, Q. Zhao, Circularly Polarized Luminescence from Organic Micro-/nano-structures, Light Sci. Appl. 2021, 10, 76. 142. E. M. Sánchez-Carnerero, A. R. Agarrabeitia, F. Moreno, B. L. Maroto, G. Muller, M. J. Ortiz, S. de la Moya, Circularly Polarized Luminescence from Simple Organic Molecules, Chem. Eur. J. 2015, 21, 13488. 143. Y.-Y. Ju, L. Chai, K. Li, J.-F. Xing, X.-H. Ma, Z.-L. Qiu, X.-J. Zhao, J. Zhu, Y.-Z. Tan, Helical Trilayer Nanographenes with Tunable Interlayer Overlaps, J. Am. Chem. Soc. 2023, 145, 2815. 144. W. Niu, Y. Fu, Z.-L. Qiu, C. J. Schürmann, S. Obermann, F. Liu, A. A. Popov, H. Komber, J. Ma, X. Feng, π-Extended Helical Multilayer Nanographenes with Layer-Dependent Chiroptical Properties, J. Am. Chem. Soc. 2023, 145, 26824. 145. W. Niu, Y. Fu, Q. Deng, Z.-L. Qiu, F. Liu, A. A. Popov, H. Komber, J. Ma, X. Feng, Enhancing Chiroptical Responses in Helical Nanographenes via Geometric Engineering of Double [7]Helicenes, Angew. Chem. Int. Ed. 2024, e202319874. 146. L. Arrico, L. Di Bari, F. Zinna, Quantifying the Overall Efficiency of Circularly Polarized Emitters, Chem. Eur. J. 2021, 27, 2920. 147. M. Buendía, J. M. Fernández-García, J. Perles, S. Filippone, N. Martín, Enantioselective Synthesis of a Two-fold Inherently Chiral Molecular Nanographene, Nat. Synth. 2024, 3, 545. 148. H. Jürgen Bestmann, W. Both, Determination of the Absolute Configuration of (+)- Pentahelicene, Angew. Chem. Int. Ed. 1972, 11, 296. 149. I. G. Stará, I. Starý, M. Tichf, J. Závada, V. Hanus, Stereochemical Dichotomy in the Stevens Rearrangement of Axially Twisted Dihydroazepinium and Dihydrothiepinium Salts. A Novel Enantioselective Synthesis of Pentahelicene, J. Am. Chem. Soc. 1994, 116, 5084. 150. K. Tanaka, H. Suzuki, H. Osuga, Non-Photochemical Route to Chiral Disubstituted [7]Thiaheterohelicenes via Biaryl- and Carbonyl-Coupling Reactions, J. Org. Chem. 1997, 62, 4465. 151. K. Nakano, Y. Hidehira, K. Takahashi, T. Hiyama, K. Nozaki, Stereospecific Synthesis of Hetero[7]helicenes by Pd-Catalyzed Double N-Arylation and Intramolecular O- Arylation, Angew. Chem. Int. Ed. 2005, 44, 7136. 152. A. Grandbois, S. K. Collins, Enantioselective Synthesis of [7]Helicene: Dramatic Effects of Olefin Additives and Aromatic Solvents in Asymmetric Olefin Metathesis, Chem. Eur. J. 2008, 14, 9323. 153. S. Chen, Z. Ge, Q. Jia, K.-P. Wang, L.-H. Gan, Z.-Q. Hu, The Preparation of Enantiopure [6]- and [7]Helicenes from Binaphthanol, Chem. Asian J. 2019, 14, 1462. 154. U. Dhawa, C. Tian, T. Wdowik, J. C. A. Oliveira, J. Hao, L. Ackermann, Enantioselective Pallada-Electrocatalyzed C−H Activation by Transient Directing Groups: Expedient Access to Helicenes, Angew. Chem. Int. Ed. 2020, 59, 13451. References 268 155. T. Ikai, K. Oki, S. Yamakawa, E. Yashima, Stereospecific Synthesis of Enantiopure [6]Helicene Containing a Seven-Membered Ring and [7]Helicene by Acid-Promoted Stepwise Alkyne Annulations of Doubly Axial-Chiral Precursors, Angew. Chem. Int. Ed. 2023, 62, e202301836. 156. J. M. Brunel, BINOL:  A Versatile Chiral Reagentm, Chem. Rev. 2005, 105, 857. 157. J. Aydin, K. S. Kumar, M. J. Sayah, O. A. Wallner, K. J. Szabó, Synthesis and Catalytic Application of Chiral 1,1‘-Bi-2-naphthol- and Biphenanthrol-Based Pincer Complexes:  Selective Allylation of Sulfonimines with Allyl Stannane and Allyl Trifluoroborate, J. Org. Chem. 2007, 72, 4689. 158. R. Kshatriya, Recent Advancement in H8–BINOL Catalyzed Asymmetric Methodologies, ACS Omega 2023, 8, 17381. 159. J. Jiménez, R. Prieto-Montero, S. Serrano, P. Stachelek, E. Rebollar, B. L. Maroto, F. Moreno, V. Martinez-Martinez, R. Pal, I. García-Moreno, S. de la Moya, BINOL Blocks as Accessible Triplet State Modulators in BODIPY Dyes, Chem. Commun., 2022, 58, 6385. 160. K. Sato, M. Hasegawa, Y. Nojima, N. Hara, T. Nishiuchi, Y. Imai, Y. Mazaki, Circularly Polarized Luminescence of a Stereogenic Curved Paraphenylene Anchoring a Chiral Binaphthyl in Solution and Solid State, Chem. Eur. J. 2021, 27,1323. 161. R. K. Dubey, M. Melle-Franco, A. Mateo-Alonso, Inducing Single-Handed Helicity in a Twisted Molecular Nanoribbon, J. Am. Chem. Soc. 2022, 144, 2765. 162. N. V. Tkachenko, S. Scheiner, Optical Stability of 1,1′-Binaphthyl Derivatives, ACS Omega 2019, 4, 6044. 163. J. Areephong, N. Ruangsupapichart, T. Thongpanchang, A Concise Synthesis of Functionalized 7-oxa-[5]-helicenes, Tetrahedron Lett. 2004, 45, 3067. 164. C. A. Guido, F. Zinna, G. Pescitelli, CPL calculations of [7]helicenes with Alleged Exceptional Emission Dissymmetry Values, J. Mater. Chem. C, 2023, 11, 10474. 165. L. Zhang, I. Song, J. Ahn, M. Han, M. Linares, M. Surin, H.-J. Zhang, J. H. Oh, J. Lin, π-Extended Perylene Diimide Double-heterohelicenes as Ambipolar Organic Semiconductors for Broadband Circularly Polarized Light Detection, Nat. Commun. 2021, 12, 142. 166. M. Cei, L. Di Bari, F. Zinna, Circularly Polarized Luminescence of Helicenes: A Data- Informed Insight, Chirality 2023, 35, 192. 167. K. Nakanishi, D. Fukatsu, K. Takaishi, T. Tsuji, K. Uenaka, K. Kuramochi, T. Kawabata, K. Tsubaki, Oligonaphthofurans: Fan-Shaped and Three-Dimensional π-Compounds, J. Am. Chem. Soc. 2014, 136, 7101. 168. S. Hossain, M. Akter, M. Karikomi, Synthesis of Oxa[9]helicene Derivatives by Cyclodehydration of 1,1′-bibenzo[c]phenanthrene-2,2′-diol Using Phosphorus Pentoxide, Tetrahedron Lett. 2021, 69, 152948. References 269 169. M. Sako, Y. Takeuchi, T. Tsujihara, J. Kodera, T. Kawano, S. Takizawa, H. Sasai, Efficient Enantioselective Synthesis of Oxahelicenes Using Redox/Acid Cooperative Catalysts, J. Am. Chem. Soc. 2016, 138, 11481. 170. W. Fudickar, T. Linker, Why Triple Bonds Protect Acenes from Oxidation and Decomposition, J. Am. Chem. Soc. 2012, 134, 15071. 171. Y. Nojima, M. Hasegawa, N. Hara, Y. Imai, Y. Mazakia, Stereogenic Cyclic Oligonaphthalenes Displaying Rring Size-dependent Handedness of Circularly Polarized Luminescence (CPL), Chem. Commun. 2019, 55, 2749. 172. Gaussian 09, Revision D.01, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman, and D. J. Fox, Gaussian, Inc., Wallingford CT, 2016. 173. (a) A. D. Becke, Density‐functional Thermochemistry. III. The Role of Exact Exchange, J. Chem. Phys. 1993, 98, 5648; b) C. Lee, W. Yang, R. G. Parr, Development of the Colle- Salvetti Correlation-energy Formula into a Functional of the Electron Density, Phys. Rev. B 1998, 37, 785; c) S. H. Vosko, L. Wilk, M. Nusair, Can. Accurate Spin-dependent Electron Liquid Correlation Energies for Local Spin Density Calculations: a Critical Analysis, J. Phys. 1980, 58, 1200. 174. S. Grimme, J. Antony, S. Ehrlich, H. Krieg, H. A Consistent and Accurate ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu, J. Chem. Phys. 2010, 132, 154104. 175. F. Weigend, R. Alhrichs, Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy, Phys. Chem. Chem. Phys. 2005, 7, 3297. 176. J. W. McIver, A. K. Komornicki, Structure of Transition States in Organic Reactions. General Theory and an Application to the Cyclobutene-butadiene Isomerization using a Semiempirical Molecular Orbital Method, J. Am. Chem. Soc. 1972, 94, 2625. 177. González, C.; Schlegel, H. B. Reaction Path Following in Mass-weighted Internal Coordinates, J. Phys. Chem. 1990, 94, 5523. 178. (a) S. Miertuš, E. Scrocco, J. Tomasi, Electrostatic Interaction of a Solute with a Continuum. A Direct Utilizaion of AB Initio Molecular Potentials for the Prevision of Solvent Effects, Chem. Phys. 1981, 55, 117; (b) J. L. Pascual-Ahuir, E. Silla, I. Tuñón, GEPOL: An Improved Description of Molecular Surfaces. III. A New Algorithm for the References 270 Computation of a Solvent-excluding Surface, J. Comp. Chem. 1994, 15, 1127; (c) V. Barone, M. Cossi, Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent Model, J. Phys. Chem. A, 1998, 102, 1995. 179. (a) M. E. Casida, Recent Developments and Applications of Modern Density Functional Theory; Elsevier: Amsterdam, 1996; Vol. 4; (b) M. E. Casida, D. P. Chong, Recent Advances in Density Functional Methods; World Scientific: Singapore, 1995; Vol. 1, p 155. 180. For a review, see: A. Dreuw, M. Head-Gordon, M. Single-Reference ab Initio Methods for the Calculation of Excited States of Large Molecules, Chem. Rev. 2005, 105, 4009. 181. Z. Chen, C. S. Wannere, C. Corminboeuf, R. Puchta, P. Schleyer, Nucleus-Independent Chemical Shifts (NICS) as an Aromaticity Criterion, Chem. Rev. 2005, 105, 3842. 182. K. Wolinski, J. F. Hilton, P. Pulay, Efficient Implementation of the Gauge-independent Atomic Orbital Method for NMR Chemical Shift Calculations, J. Am. Chem. Soc. 1990, 112, 8251. 183. (a) Becke, A. D. Density‐functional Thermochemistry. III. The Role of Exact Exchange, J. Chem. Phys. 1993, 98, 5648; (b) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti Correlation-energy Formula into a Functional of the Electron Density, Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785; (c) Vosko, S. H.; Wilk, L.; Nusair, M. Can. Accurate Spin-dependent Electron Liquid Correlation Energies for Local Spin Density Calculations: a Critical Analysis, J. Phys. 1980, 58, 1200. 184. Johnson, E. R.; Keinan, S.; Mori-Sánchez, P.; Contreras-García, J.; Cohen, A. J.; Yang, W. Revealing Noncovalent Interactions, J. Am. Chem. Soc. 2010, 132, 6498. Tesis Patricia Izquierdo García PORTADA TABLE OF CONTENTS SUMMARY RESUMEN INTRODUCTION CHAPTER 1 CHAPTER 2 CHAPTER 3 CHAPTER 4 EXPERIMENTAL SECTION APPENDIX 1. Chapter 1 APPENDIX 2. Chapter 2 APPENDIX 3. Chapter 3 APPENDIX 4. Chapter 4 APPENDIX 5. Spectra REFERENCES