Inducing Open-shell Character in Porphyrins through Surface-as- sisted Phenalenyl π-Extension

Qiang Sun,⊥,¶, Luis M. Mateo,†,‡,¶, Roberto Robles,‖,¶ Pascal Ruffieux,⊥ Nicolas Lorente,*,‖,º Gio- vanni Bottari,*,†,‡,§ Tomás Torres*,†,‡,§ and Roman Fasel*,⊥,+

⊥nanotech@surfaces Laboratory, Empa-Swiss Federal Laboratories for Materials Science and Technology, 8600 Dübendorf, Switzerland †Departamento de Química Orgánica, Universidad Autónoma de Madrid, 28049 Madrid, Spain ‡IMDEA-Nanociencia, Campus de Cantoblanco, 28049 Madrid, Spain ‖Centro de Física de Materiales CFM/MPC (CSIC-UPV/EHU), Paseo de Manuel de Lardizabal 5, 20018 Donostia-San Sebastián, Spain º Donostia International Physics Center (DIPC), 20018 Donostia-San Sebastián, Spain §Institute for Advanced Research in Chemical Sciences (IAdChem), Universidad Autónoma de Madrid, 28049 Ma- drid, Spain +Department of Chemistry and Biochemistry, University of Bern, 3012 Bern, Switzerland ABSTRACT: Organic open-shell compounds are extraordinarily attractive materials for their use in molecular spintronics thanks Charge-transfer Inherent induced open-shell open-shell to their long spin-relaxation times and structural flexibility. Por- doublet character triplet character phyrins (Pors) have widely been used as molecular platforms to craft persistent open-shell structures through solution-based re- PorA4 S = 1/2 PorA2 S = 1 dox chemistry. However, very few examples of inherently open- shell Pors have been reported, which are typically obtained through the fusion of non-Kekulé polyaromatic hydrocarbon moi- J = 19.4 mV eties to the Por core. The inherent instability and low solubility of these radical species, however, requires the use of bulky substitu- ents and multi-step synthetic approaches. On-surface synthesis has emerged as a powerful tool to overcome such limitations, giv- ing access to structures that cannot be obtained through classical methods. Herein, we present a simple and straightforward method for the on-surface synthesis of phenalenyl-fused Pors using readily available molecular precursors. In a systematic study, we examine the structural and electronic properties of three surface-supported Pors, bearing zero, two (PorA2) and four (PorA4) meso-fused phenalenyl moieties. Through atomically resolved real-space imaging by scanning probe micros- copy and high-resolution scanning tunneling spectroscopy combined with density functional theory (DFT) calculations, we unambiguously demonstrate a triplet ground state for PorA2 and a charge transfer induced open-shell character for the intrinsically closed-shell PorA4.

INTRODUCTION nature of the peripheral substituents.4 Among such struc- Porphyrins (Pors), tetrapyrrolic macrocycles with a planar tural modifications, the π-extension of the Por core is par- structure and an aromatic circuit of 18 π-electrons, are het- ticularly appealing, since it leads to compounds with a erocycles showing high thermal stability,1 rich coordination strong near-infrared (NIR) absorption, stemming from a chemistry,2 tunable redox features,3 and excellent photo- small highest occupied molecular orbital (HOMO)-lowest 12 physical properties.4 Their unique features make these unoccupied molecular orbital (LUMO) gap. Such π-ex- chromophores prime candidates for a wide range of appli- tended Pors can be obtained either by attaching other aro- 9,13-16 cations ranging from materials science,5 catalysis,6 photo- matic units to the macrocycle via conjugative linkers or voltaics,7-10 or medicine,11 to mention a few. through the meso,β-fusion of aromatic units via intramo- lecular oxidative coupling.17-24 Using the latter synthetic Remarkably, many of these properties can be precisely strategy, a wide range of π-extended Pors has been re- tuned through chemical modifications, such as the com- ported, showing, in some cases, absorptions reaching far plexation of metal ions2 or the modification of number and into the NIR region (i.e., 1417 nm) and optical HOMO- LUMO gaps as low as 0.61 eV.25 However, the vast majority strated by DFT calculations and resonance structure analy- 26 of these systems presents a closed-shell character, while sis. Further proof of the open-shell character of PorA2 is only few examples of open-shell analogues have been re- provided by the coexistence of singly and doubly β-hydro- 27,28 ported. Organic open-shell structures are extraordinar- genated Pors with the intact PorA2. Furthermore, we have ily attractive materials in spintronics, presenting longer explored the electronic properties of all three fabricated spin-relaxation times and easier tunability than inorganic molecules by STS. For PorA2, an open-shell triplet ground materials.27,29-31 Along this line, Osuka and co-workers re- state is experimentally demonstrated by inelastic spin exci- ported the preparation of several inherent and persistent tation and Kondo scattering at low energies. For PorA4, Por-based radicals using rational molecular design and, in DFT calculations and resonance structure analysis both most examples, redox chemistry.28 These radicals are either predict a closed-shell electronic structure. Interestingly, charged (anions/cations) or neutral with the unpaired elec- charge transfer to the underlying surface gives rise to in- tron poorly delocalized over the macrocycle. duced open-shell character to the intrinsically closed-shell Over the past years, the preparation and study of π-ex- PorA4, as reflected by a characteristic Kondo resonance. tended Pors with highly delocalized radicals has, therefore, b) gathered significant attention. For this purpose, an inter- esting yet poorly explored strategy consists in the fusion of inherently radical polyaromatic hydrocarbons (PAHs) to the macrocycle.32-34 Among these PAHs, non-Kekulé struc- tures like phenalenyl – its smallest member – have been widely used as a building block to generate radicaloid and biradicaloid structures with remarkable thermodynamic d) stability.27,29,35-40 However, the synthetic access to phenalenyl-based structures is often hampered by multi- step synthesis, low yields, and poorly soluble intermediates and products. Moreover, the inherently low HOMO-LUMO gap of phenalenyl leads to stability issues, such that the characterization of these species needs to be carried out in degassed solutions or through in-situ generation of the rad- ical species.32 On-surface synthesis stands as a promising tool to prepare open-shell PAH systems, offering significant advantages over the more classical “wet” synthesis. The employed ultra- high vacuum (UHV) conditions and the atomically clean, f) catalytic surfaces used, provide the ideal playground for both synthesis and characterization of structures that are hardly accessible through solution-based methods.41-51 In- deed, recently, elusive PAHs or nanographenes with open- shell characters such as Clar's goblet and triangulenes have been successfully realized through on-surface synthe- sis.48,52-55 The on-surface chemistry of Pors and related por- phyrinoids has emerged in the past decades, focusing mainly on model systems like porphines,56,57 tetra- phenylPors,58-60 and derivatives thereof.61 On-surface fabri- cation of π-extended Pors by fusing the macrocycle with PAHs has only recently been reported.57,62,63 However, the few reported examples involve hybrids of Por with nanog- raphenes, not focusing on the possible open-shell character of the π-system. Figure 1. Molecular structures of a) PorA0, c) PorA2, and e) Herein, we present the on-surface synthesis and character- PorA4. In c,e), turquoise hexagons identify Clar’s sextets, ization of three Zn(II)Pors, namely PorA0, PorA2, and whereas in a,e), red dashed lines indicate a 18 π-electron cir- PorA4, with zero, two, and four meso,β,β triply-fused cuit. HOMA and bond length analysis of b) PorA0, d) PorA2, phenalenyl moieties, respectively. While PorA0 was pre- and f) PorA4. The scale bars for bond length and HOMA apply pared by “wet” synthesis, the two- and four-fold to b), d) and f). Zn-N bonds do not correlate with the color phenalenyl-fused Por derivatives PorA2 and PorA4 were scale and have been colored in orange. fabricated through a surface-assisted cyclodehydrogena- RESULTS AND DISCUSSION tion reaction64 from meso-2,6-dimethylphenyl (dmp) sub- Synthesis of PorA0, and precursors of PorA2 and PorA4 stituted precursors Por(dmp)2 and Por(dmp)4, respec- tively. This study reveals the electronic impact of the D4h-symmetric H2PorA0 and H2Por(dmp)4 were prepared phenalenyl substituents fused to the Por macrocycle, lead- by condensing pyrrole with formaldehyde or 2,6-dime- ing to the open-shell character for PorA2, which is demon- thylbenzalehyde, respectively, under classical Lindsey con- ditions (Scheme 1a,c).65 While the condensation reaction

involving 2,6-dimethylbenzaldehyde afforded which are tentatively identified as covalently-linked dimers H2Por(dmp)4 in 15% yield, the low reactivity and solubility and trimers (Figure 2d,g). High-resolution STM images of formaldehyde limited the yield of H2PorA0 (i.e., 1.5%) (Figure 2b,e,h) reveal fully planarized structures with the with a considerable amount of polypyrrolic tar obtained as expected symmetry (i.e., D4h for PorA0 and PorA4, and D2h side-product. D2h-symmetric H2Por(dmp)2 was prepared for PorA2), suggesting the successful formation of the tar- in 27% yield by condensing meso-H-dipyrromethane with get structures. Finally, the molecular structures were unam- 2,6-dimethylbenzaldehyde under standard Lindsey condi- biguously confirmed by bond-resolved nc-AFM imaging tions (Scheme 1b). Subsequent metalation of H2PorA0, (Figure 2c,f,i). H Por(dmp) , and H Por(dmp) with Zn(OAc) in reflux- 2 2 2 4 2 In a first step, the potential open-shell character of PorA0, ing THF afforded the corresponding zinc metalated deriva- PorA2, and PorA4 was evaluated based on their resonance tives PorA , Por(dmp) , and Por(dmp) , respectively, in 0 2 4 structures. PorA0 presents a conjugated circuit of 18 π-elec- excellent yields. PorA0, Por(dmp)2, and Por(dmp)4, as trons, fulfilling Hückel's rule (Figure 1a), and its predicted well as their precursors, were fully characterized by a wide electronic character is closed-shell. However, the fusion of range of spectroscopic and spectrometric techniques (see two and four phenalenyls to the Por core leading to PorA2 Supporting Information). and PorA4, respectively, gives rise to non-Kekulé resonance Scheme 1. Synthetic route to the preparation of a) structures (Figure 1c,e and Figures S4.1 and S4.2). In the a PorA0, b) PorA2, and c) PorA4. case of PorA2, several resonance structures, all of which are non-Kekulé, can be drawn (Figure S4.1). Among those, the ones which show the maximum number of Clar’s sextets (i.e., 2) and minimum number of unpaired electrons (i.e., 2) ‒ as the one drawn in Figure 1c (vide infra) ‒ can be ex- pected to contribute most to the ground state electronic structure. It is worth to notice that none of the relevant res- onance structures preserves the inner Por 18 π-electron ar- omatic circuit. For PorA4, a closed-shell Kekulé structure with four Clar's sextets on the phenalenyls and an 18 π-elec- tron circuit on the central macrocycle can be drawn (Figure 1e). In addition, several non-Kekulé structures can be pro- posed (Figure S4.2), one of which minimizes the number of radicals (i.e., 2) while maintaining the four Clar's sextets (Figure S4.2b). To obtain further insights, harmonic oscilla- tor model of aromaticity (HOMA)66,67 analysis has been performed on DFT optimized PorA2 and PorA4 in the gas phase. These studies show excellent agreement with the proposed resonance structures, assigning high and low de- gree of benzenoid character to the outermost six-mem- bered rings and the pyrrolic units, respectively (Figure 1d,f). Furthermore, gas-phase DFT calculations at the B3LYP level show that, for PorA2, the open-shell electronic config- uration is 0.46 eV more stable than the closed-shell one. On the other hand, in the case of PorA4, the closed-shell elec- tronic configuration is 0.21 eV more stable than the open- shell one by theory.

Particular attention was therefore paid to PorA2, aimed at confirming its predicted open-shell character. Surprisingly, after a close inspection of the reaction products by taking a a Reagents and conditions: i) 1) BF3O(Et)2, r.t., 2) DDQ, reflux, number of high-resolution nc-AFM images, we found that 3) NEt3, r.t. ii) Zn(OAc)2, THF, reflux; DDQ = 2,3-dichloro-5,6- there were two other species besides PorA2 (Figure 3, panel dicyano-1,4-benzoquinone. iii) which were almost indiscernible from normal topo-

On-surface Synthesis and Characterization of PorA0, graphic STM images (Figure 3, panel ii). One species fea- PorA2, and PorA4. tures two bright protrusions at opposite β-sites of the mac- rocycle in nc-AFM images (Figure 3a, panel iii), while an- PorA0 was deposited on Au(111) at room temperature and examined by STM imaging after cool down to 4.5 K, yield- other only features one protrusion at a β-site (Figure 3b, ing samples of high purity and homogeneity (Figure 2a). panel iii). Since constant-height nc-AFM imaging is very On the other hand, samples with deposited precursors sensitive to the apparent height of the underlying struc- tures, each bright protrusion suggests a doubly - Por(dmp)2 and Por(dmp)4 were annealed above 300 ºC to 3 promote the surface-assisted cyclodehydrogenation be- ated sp hybridized carbon (CH2) as also observed in other open-shell nanographene systems.48,49 Remarkably, these tween the methyl groups and the macrocycle, giving PorA2 63 hydrogenated β positions correspond with the sites of the and PorA4, respectively. The corresponding large-scale unpaired electrons in the proposed resonance structure of STM images of PorA2 and PorA4 confirm the presence of individual molecules, but also reveal several side-products, PorA2 (Figure 1c).

Figure 2. a) Large-scale STM image of PorA0 (chemical structure in the inset) as-deposited on Au(111) at room temperature (Vs = ‒ 0.5 V, It = 50 pA). b) Close-up STM (Vs = ‒0.6 V, It = 400 pA) and c) constant-height nc-AFM (Vs = ‒0.001 V) images of PorA0. d) Large-scale STM image of PorA2 (chemical structure in the inset) obtained upon annealing of Por(dmp)2 at 300 °C on Au(111) (Vs = ‒0.2 V, It = 50 pA). e) Close-up STM (Vs = ‒0.06 V, It = 150 pA) and f) constant-height nc-AFM (Vs = ‒0.005 V) images of PorA2. g) Large-scale STM image of PorA4 (chemical structure in the inset) obtained upon annealing of Por(dmp)4 at 330 °C on Au(111) (Vs = ‒0.2 V, It = 50 pA). h) Close-up STM (Vs = ‒0.1 V, It = 150 pA) and i) constant-height nc-AFM (Vs = ‒0.005 V) images of PorA4. STM induced atomic manipulation can be employed to se- features in the differential conductance (dI/dV) spectros- 3 lectively break one C-H bond and transform C(sp )H2 into copy revealing inelastic spin excitations and/or a Kondo 2 68 C(sp )H, as the C-H bond-dissociation energy is lower in resonance. Thus, dI/dV spectra were taken on PorA2 with 48,49 CH2 than in CH. To further elucidate the chemical a focus on the low bias range (Figure 3d). Interestingly, we structures of the hydrogen passivated species, we applied a observed characteristic conductance steps symmetrically corresponding STM manipulation protocol. The tip was in- located around the Fermi level, along with a zero-bias peak. itially positioned above a bright protrusion (CH2) of a di- The step-like conductance increase can be associated with hydrogenated PorA2(H2) with a typical STM setpoint of V the inelastic excitation between the ferromagnetic (triplet) = ‒0.1 V and I = 10 pA (Figure 3a). The tip was thereafter ground state and an antiferromagnetic (singlet) excited retracted by 4 Å to limit the tunneling current at the open- state, while the zero-bias peak is usually related to the feedback loop. Then the voltage bias was gradually ramped many-body Kondo state arising from the screening of a lo- up to 3.4 eV until an abrupt change of the tunneling current calized spin by itinerant electrons from the metal surface.69 occurred, indicating a manipulation event. The resulting However, unlike spin excitation spectra for organic systems structure was imaged by both high-resolution STM and nc- with singlet ground states,52,70 the coexistence of conduct- AFM, revealing the formation of the mono-hydrogenated ance steps and a zero-bias peak points toward a triplet S = 1 71 PorA2(H) species (Figure 3b). Repeating the same protocol, ground state. The low-energy dI/dV spectrum can be PorA2(H) can be further transformed into PorA2 (Figure nicely fit by a dynamical scattering model developed by M. 3c). The success rate of such atomic manipulation is as high Ternes,68 in which we constructed a S = 1 system with two as 90% without disrupting the molecular structure. These ferromagnetically exchange-coupled spins. An exchange manipulation experiments suggest that PorA2(H) and coupling energy J = 19.4 meV was derived from the fit. Our PorA2(H2) are derived from initially formed PorA2 through DFT calculations on the surface also predict a ferromag- its passivation with H2 from the cyclodehydrogenation pro- netic ground state and an antiferromagnetic state with a cess. DFT calculations on the surface show that passivating higher energy, consistent with the experimental observa- PorA2 is favorable by 0.79 eV per H, in agreement with the tion. Moreover, we found that inclusion of a non-zero experimental observations. The presence of PorA2(H) and Kondo scattering parameter Jρs is crucial to reproduce both PorA2(H2) implies a high reactivity of PorA2, indicating a the cusp above the excitation energy and the zero-bias peak certain radical character, in line with the predicted open- in the spectrum, with a value of ‒0.14 determined from the shell structure (Figure 1b). fit. In Figure 3e, we show the spin excitation maps of PorA2 An important and direct fingerprint of molecules with obtained by constant height STS at ± 22 mV, which agree open-shell character is the presence of low-energy spectral very well with the predicted spin density map. The effect of the substrate will be discussed by our detailed DFT calcula- tions in the following context.

Figure 3. a) i): Chemical structure of PorA2(H2); ii): high-resolution STM (Vs = ‒0.04 V, It = 100 pA) and iii) nc-AFM (Vs = ‒0.005 V) images where the doubly hydrogenated sites are indicated by arrows. b) i): Chemical structure of PorA2(H); ii) its high-resolution STM (Vs = ‒0.06 V, It = 200 pA) and iii) nc-AFM (Vs = ‒0.005 V) images where the doubly hydrogenated site is indicated by an arrow. c) i): Chemical structure of PorA2; ii): its high-resolution STM (Vs = ‒0.06 V, It = 200 pA) and iii) nc-AFM (Vs = ‒0.005 V) images. The stepwise transformation of a) PorA2(H2) to b) PorA2(H) to c) PorA2 can be achieved by STM tip induced hydrogen abstraction. Scale bar: 5 Å. d) dI/dV spectrum acquired over PorA2 at the position marked in c) (Vmod = 0.8 mV). The data is shown by blue markers, and the solid red line is a fit based on a dynamical scattering model developed by M. Ternes.68 DFT calculated spin density maps for the ferromagnetic and antiferromagnetic ground states are displayed in the inset, where red and yellow densities denote the spin-up and spin-down contributions, respectively. e) Constant-height STS maps taken at bias voltages of ‒22 mV and 22 mV (Vmod = 4 mV). Scale bar: 5 Å.

The electronic properties of PorA0, PorA2, and PorA4 were dI/dV spectrum exhibits three broad peaks in the probed further investigated by wider bias range differential con- energy range (Figure 4e). Spin-polarized DFT calculations ductance dI/dV spectroscopy to probe their local density of were performed to elucidate the electronic structure of states (LDOS). The dI/dV spectrum of PorA0 reveals a wide PorA2 on Au(111) (Figure S2.2). They reveal a rather compli- energy gap around the Fermi level which is enclosed by two cated energy level renormalization for this open-shell mol- broad peaks beyond ±1V (Figure 4a). The energy gap on ecule (Figure 4g). The DOS projected on the frontier mo- Au(111), around 2.3 eV derived from the onsets of the peaks, lecular orbitals shows that the gas-phase HOMO of PorA2 is comparable to the one of tetraphenyl-Por on Au(111),72 gets empty for the spin-up electrons upon adsorption on and the corresponding negative and positive ion reso- Au(111), which is also reflected in the calculated charge nances can accordingly be assigned to the Por HOMO and transfer of −0.73 electrons. We note that a direct compari- LUMO, respectively. dI/dV maps were acquired at the onset son between the dI/dV maps of PorA2 acquired at different energies of the two peaks (Figure 4b), showing the spatially bias voltages with the DFT computed dI/dV maps is diffi- resolved HOMO (‒1.0 V) and LUMO (1.3 V) orbitals. DFT cult due to the underestimation of the HOMO-LUMO gap calculations of the PorA0 on Au(111) were performed to give inherent to DFT. In order to ease the comparison, we have further insights (Figure S2.1). Both the experimental dI/dV used the cation of PorA2 (with 1 electron being removed) in spectrum and maps are in good agreement with the DFT the gas phase as a model system of the surface supported computed projected density of states (PDOS) of the sur- molecule. Calculations of the gas phase cation at the B3LYP face-supported PorA0 (Figure 4c) and its corresponding level of theory combined with the results on the surface ob- simulated dI/dV maps (Figure 4d), respectively. The fusion tained at the PBE level (see discussions in Figure S2.2 and of two phenalenyls to the Por core in PorA2 significantly al- S2.3) allow us to assign the peak at 0.6 V in Figure 4e to the ters the molecular energy level structure, and the resulting empty molecular spin-up HOMO, while the peak starting

from 1.3 V is assigned to the spin-down LUMO and agrees with the HOMO getting empty in the DOS projected LUMO+1 (Figure 4f-h) and the broad peak at the negative on molecular orbitals (Figure S2.4 and Figure 4k). We can side from ‒0.4 V is assigned to the spin-up HOMO-1. It is therefore assign the broad feature in the dI/dV spectrum to important to note that, despite the charge transfer of PorA2 the emptied spin-down HOMO, the doubly degenerate to the Au(111) substrate, a ferromagnetic ground state is also LUMO, and the LUMO+1. The sharp peak at 1.9 V is as- predicted by DFT (Figure S2.3). signed to the LUMO+2. This assignment of molecular or- bitals is corroborated by the excellent match between ex- Turning to PorA4, its dI/dV spectrum shows a broad density of states from 0.5 V to 1.2 V and a sharp peak around 1.9 V perimental and theoretical maps (Figure 4j and 4l, respec- (Figure 4i). Our DFT calculations indicate a charge transfer tively), which further confirms the DFT prediction of a siz- of −1.11 electrons from the molecule to the surface, which able electron transfer from the Por to the substrate.

Figure 4. a,e,i) Wide bias range STS spectrum of PorA0, PorA2 and PorA4, respectively. The corresponding STM images and the acquisition position of the STS spectra are shown in the inset; b,f,j) constant-current STS maps recorded at the energies of the prominent peaks; c,g,k) DFT computed PDOS (black curve) on the surface-supported PorA0, PorA2 and PorA4, respectively, where the colored areas indicate the PDOS on its corresponding gas-phase molecular orbitals (gas phase molecular orbitals indicated in Figure S2); d) simulated dI/dV maps matching the experimental ones in (b); h) simulated dI/dV maps matching the experimental ones in (f); l) simulated dI/dV maps matching the experimental ones in (j). Scale bars in b,d,f,h,j,l): 5 Å.

0 PorA4 were obtained by surface-assisted cyclodehydro- a) .

b) 16 genation from appropriate meso-(2,6-dimethylphenyl)- functionalized Por precursors.

I (

nA The potential open-shell character of these three Pors was

) evaluated, firstly, by consideration of their resonance structures and by gas-phase DFT calculations. In the case of PorA0, only Kekulé resonance structures can be drawn,

0 suggesting a closed-shell character, which is confirmed by DFT. On the other hand, DFT calculations on PorA2 pre- c) Exp. data d) HWHM from Frota fit dict an open-shell character, which is in line with the non- Frota fit Fermi-liquid fit: 17.0 K Kekulé resonance structures that can be proposed for this TKondo = 78.2 ± 1.7 K Por. In the case of PorA4, DFT calculations and resonance 14.3 K 12.8 K structure analysis both predict a closed-shell character, while some reasonable non-Kekulé resonance structures 10.0 K can also be proposed. 7.9 K The three surface-supported Pors were characterized by 4.5 K high-resolution STM and atomically resolved nc-AFM im- aging. Interestingly, in the case of PorA2, the presence of singly and doubly β-hydrogenated Por species was also ob- served, further supporting the hypothesis of a marked Figure 5. a) Low-bias STM topography (Vs = –0.05 V, It = 450 open-shell character. Moreover, the doubly hydrogenated pA) and b) constant-height STM image (Vs = 5 mV) of PorA4 Por could be converted into the singly hydrogenated spe- on Au(111). Scale bar: 5 Å. c) Temperature evolution of the STS cies and the latter into pristine PorA2 by STM atomic ma- spectra for PorA with the experimental data fit by the Frota 4 nipulation. Unequivocal evidence for an open-shell char- function. The acquisition position of STS spectra is marked in acter of PorA2 was obtained from low-bias STS, where we a) (Vmod = 0.8 mV). d) Extracted half width at half maximum (HWHM) of the Kondo resonance as a function of tempera- found coexistence of inelastic excitations and a zero-bias ture, which is fit using the Fermi-liquid model to determine peak, suggesting a triplet ground state. Additionally, the Kondo temperature. charge transfer to the surface gives rise to induced open- shell character to the intrinsically closed-shell PorA4, The charging of a closed-shell molecule as a consequence which presents a characteristic Kondo resonance. Our re- of molecule-to-surface charge transfer could give rise to un- sults evidence a facile strategy to the formation of open- paired electrons on the molecule and induce a magnetic shell Por molecules, which are promising for future molec- moment.73 Indeed, we observed a pronounced zero-bias ular electronics and spintronics. resonance in low-energy dI/dV spectra acquired on PorA4 (Figure 5c). The resonance can be fit by a Frota function de- scribing a Kondo resonance, from which the resonance ASSOCIATED CONTENT width (half width at half maximum, HWHM) can be ab- The Supporting Information is available free of charge on the stracted. The HWHM of the resonance exhibits a charac- ACS Publications website at DOI: teristic temperature dependent broadening following the Fermi-liquid model, evidencing a Kondo resonance with a AUTHOR INFORMATION Kondo temperature T = 78 K extracted from the Fermi-liq- K Corresponding Authors uid fit (Figure 5d). This unambiguously evidences the pres- ence of an unpaired electron with S=1/2 on the surface sup- *[email protected] *[email protected] ported PorA4 molecule. The charge-transfer induced mag- *[email protected] netic properties of the surface supported PorA agree with 4 *[email protected] the DFT calculations that indicate a charge transfer of ‒1.11 electrons and a magnetic moment of 0.69 µB, compatible with a S=1/2 ground state. Moreover, constant-height cur- Author Contributions rent maps around the Kondo peak (V = 5 mV), approximat- ¶Q.S., L.M.M. and R.R. contributed equally to this work. ing the Kondo map of PorA4 on Au(111) (Figure 5b) show a good agreement with the corresponding DFT computed Notes spin density map (Figure S2.5). The authors declare no competing financial interest. ACKNOWLEDGMENT CONCLUSION This work was supported by the Swiss National Science Foun- We have reported an in-depth study on the effect of dation under Grant No. 200020_182015. Financial support from phenalenyl -extension of Pors as an efficient way to in- Spanish MICINN (CTQ2017-85393-P) and ERA-NET/European duce open-shell character in Pors. For this purpose, three Commission/MINECO (UNIQUE, SOLAR-ERA.NET Cofund 2 N° 008/ PCI2019-111889-2) are acknowledged. IMDEA Nano- Pors presenting zero (i.e., PorA0), two (i.e., PorA2), or four ciencia acknowledges support from the “Severo Ochoa” Pro- (i.e., PorA ) phenalenyl moieties fused to the macrocycle 4 gramme for Centres of Excellence in R&D (MINECO, Grant were investigated. While PorA was prepared by solution 0 SEV2016-0686). R.R. and N.L. are grateful for funding from the chemistry and deposited as such on Au(111), PorA2 and EU-FET Open H2020 Mechanics with Molecules project (grant

766864). Q.S. thanks Kristjan Eimre for helping in HOMA (18) Richeter, S.; Jeandon, C.; Kyritsakas, N.; Ruppert, R.; Callot, analysis. H. J., Preparation of six isomeric bis-acylporphyrins with chromo- phores reaching the near-infrared via intramolecular friedel−crafts REFERENCES reaction, J. Org. Chem. 2003, 68, 9200-9208. (19) Gill, H. S.; Harmjanz, M.; Santamaría, J.; Finger, I.; Scott, M. (1) Liu, Z.; Yasseri, A. A.; Lindsey, J. S.; Bocian, D. F., Molecular J., Facile oxidative rearrangement of dispiro-porphodimethenes to memories that survive silicon device processing and real-world op- nonplanar and sheetlike porphyrins with intense absorptions in eration, Science 2003, 302, 1543-1545. the near-IR region, Angew. Chem. Int. Ed. 2004, 43, 485-490. (2) Suijkerbuijk, B. M. J. M.; Klein Gebbink, R. J. M., Merging (20) Kurotobi, K.; Kim, K. S.; Noh, S. B.; Kim, D.; Osuka, A., A porphyrins with organometallics: Synthesis and applications, An- quadruply -fused porphyrin with intense near-ir absorp- gew. Chem. Int. Ed. 2008, 47, 7396-7421. tion and a large two-photon absorption cross section, Angew. (3) Boulas, P. L.; Gómez-Kaifer, M.; Echegoyen, L., Electrochem- Chem. Int. Ed. 2006, 45, 3944-3947. istry of supramolecular systems, Angew. Chem. Int. Ed. 1998, 37, (21) Tokuji, S.; Takahashi, Y.; Shinmori, H.; Shinokubo, H.; 216-247. Osuka, A., Synthesis of a pyridine-fused porphyrinoid: Oxopyrido- (4) Mandal, A. K.; Taniguchi, M.; Diers, J. R.; Niedzwiedzki, D. chlorin, Chem. Commun. 2009, 1028-1030. M.; Kirmaier, C.; Lindsey, J. S.; Bocian, D. F.; Holten, D., Photo- (22) Jiao, C.; Huang, K.-W.; Guan, Z.; Xu, Q.-H.; Wu, J., N-annu- physical properties and electronic structure of porphyrins bearing lated fused porphyrins with enhanced near-ir absorption zero to four meso-phenyl substituents: New insights into seem- and emission, Org. Lett. 2010, 12, 4046-4049. ingly well understood tetrapyrroles, J. Phys. Chem. A 2016, 120, (23) Diev, V. V.; Hanson, K.; Zimmerman, J. D.; Forrest, S. R.; 9719-9731. Thompson, M. E., Fused –diporphyrins: Shifting near-in- (5) Jurow, M.; Schuckman, A. E.; Batteas, J. D.; Drain, C. M., Por- frared absorption to 1.5 μm and beyond, Angew. Chem. Int. Ed. phyrins as molecular electronic components of functional devices, 2010, 49, 5523-5526. Coord. Chem. Rev. 2010, 254, 2297-2310. (24) Tanaka, T.; Osuka, A., Conjugated porphyrin arrays: Syn- (6) Beyene, B. B.; Hung, C.-H., Recent progress on metallopor- thesis, properties and applications for functional materials, Chem. phyrin-based hydrogen evolution catalysis, Coord. Chem. Rev. Soc. Rev. 2015, 44, 943-969. 2020, 410, 213234. (25) Davis, N. K. S.; Thompson, A. L.; Anderson, H. L., A por- (7) Hasobe, T., Porphyrin-based supramolecular nanoarchitec- phyrin fused to four , J. Am. Chem. Soc. 2011, 133, 30- tures for solar energy conversion, J. Phys. Chem. Lett. 2013, 4, 1771- 31. 1780. (26) Mori, H.; Tanaka, T.; Osuka, A., Fused porphyrinoids as (8) Urbani, M.; Grätzel, M.; Nazeeruddin, M. K.; Torres, T., promising near-infrared absorbing dyes, J. Mater. Chem. C 2013, 1, Meso-substituted porphyrins for dye-sensitized solar cells, Chem. 2500-2519. Rev. 2014, 114, 12330-12396. (27) Sun, Z.; Ye, Q.; Chi, C.; Wu, J., Low band gap polycyclic hy- (9) Mathew, S.; Yella, A.; Gao, P.; Humphry-Baker, R.; Curchod, drocarbons: From closed-shell near infrared dyes and semicon- B. F. E.; Ashari-Astani, N.; Tavernelli, I.; Rothlisberger, U.; Nazee- ductors to open-shell radicals, Chem. Soc. Rev. 2012, 41, 7857-7889. ruddin, M. K.; Grätzel, M., Dye-sensitized solar cells with 13% effi- (28) Shimizu, D.; Osuka, A., Porphyrinoids as a platform of sta- ciency achieved through the molecular engineering of porphyrin ble radicals, Chem. Sci. 2018, 9, 1408-1423. sensitizers, Nat. Chem. 2014, 6, 242-247. (29) Morita, Y.; Suzuki, S.; Sato, K.; Takui, T., Synthetic organic (10) Wang, C.-L.; Hu, J.-Y.; Wu, C.-H.; Kuo, H.-H.; Chang, Y.-C.; spin chemistry for structurally well-defined open-shell graphene Lan, Z.-J.; Wu, H.-P.; Wei-Guang Diau, E.; Lin, C.-Y., Highly effi- fragments, Nat. Chem. 2011, 3, 197-204. cient porphyrin-sensitized solar cells with enhanced light harvest- (30) Feringa, B. L. (ed.) Molecular switches; Wiley-VCH Verlag ing ability beyond 800 nm and efficiency exceeding 10%, Energy GmbH, 2001. Environ. Sci. 2014, 7, 1392-1396. (31) Coronado, E.; Epsetin, A. J., Molecular spintronics and (11) Rajora, M. A.; Lou, J. W. H.; Zheng, G., Advancing porphy- quantum computing, J. Mater. Chem. 2009, 19, 1670. rin's biomedical utility via supramolecular chemistry, Chem. Soc. (32) Zeng, W.; Lee, S.; Son, M.; Ishida, M.; Furukawa, K.; Hu, P.; Rev. 2017, 46, 6433-6469. Sun, Z.; Kim, D.; Wu, J., Phenalenyl-fused porphyrins with differ- (12) Anderson, H. L., Building molecular wires from the colours ent ground states, Chem. Sci. 2015, 6, 2427-2433. of life: Conjugated porphyrin oligomers, Chem. Commun. 1999, (33) Diev, V. V.; Femia, D.; Zhong, Q.; Djurovich, P. I.; Haiges, R.; 2323-2330. Thompson, M. E., A quinoidal bis-phenalenyl-fused porphyrin (13) Screen, T. E. O.; Blake, I. M.; Rees, L. H.; Clegg, W.; Borwick, with supramolecular organization and broad near-infrared ab- S. J.; Anderson, H. L., Making conjugated connections to porphy- sorption, Chem. Commun. 2016, 52, 1949-1952. rins: A comparison of alkyne, alkene, imine and azo links, J. Chem. (34) Umetani, M.; Naoda, K.; Tanaka, T.; Lee, S.-K.; Oh, J.; Kim, Soc., Perkin Trans. 1 2002, 320-329. D.; Osuka, A., Synthesis of di-peri-dinaphthoporphyrins by ptcl2- (14) Chang, J.-C.; Ma, C.-J.; Lee, G.-H.; Peng, S.-M.; Yeh, C.-Y., mediated cyclization of quinodimethane-type porphyrins, Angew. Porphyrin–triarylamine conjugates: Strong electronic communi- Chem. Int. Ed. 2016, 55, 6305-6309. cation between triarylamine redox centers via the porphyrin dica- (35) Ohashi, K.; Kubo, T.; Masui, T.; Yamamoto, K.; Nakasuji, K.; tion, Dalton T 2005, 1504-1508. Takui, T.; Kai, Y.; Murata, I., 4,8,12,16-tetra-tert-butyl-s-inda- (15) Zhou, Y.; Ngo, K. T.; Zhang, B.; Feng, Y.; Rochford, J., Syn- ceno[1,2,3-cd:5,6,7-c‘d‘]diphenalene: A four-stage amphoteric thesis, electronic and photophysical characterization of π-conju- redox system, J. Am. Chem. Soc. 1998, 120, 2018-2027. gated meso-ferrocenyl-porphyrin fluorescent redox switches, Or- (36) Goto, K.; Kubo, T.; Yamamoto, K.; Nakasuji, K.; Sato, K.; ganometallics 2014, 33, 7078-7090. Shiomi, D.; Takui, T.; Kubota, M.; Kobayashi, T.; Yakusi, K.; (16) Hadmojo, W. T.; Lee, U.-H.; Yim, D.; Kim, H. W.; Jang, W.- Ouyang, J., A stable neutral hydrocarbon radical: Synthesis, crys- D.; Yoon, S. C.; Jung, I. H.; Jang, S.-Y., High-performance near-in- tal structure, and physical properties of 2,5,8-tri-tert-butyl- frared absorbing n-type porphyrin acceptor for organic solar cells, phenalenyl, J. Am. Chem. Soc. 1999, 121, 1619-1620. ACS Appl. Mater. Interfaces 2018, 10, 41344-41349. (37) Kubo, T.; Sakamoto, M.; Akabane, M.; Fujiwara, Y.; Yama- (17) Lewtak, J. P.; Gryko, D. T., Synthesis of π-extended porphy- moto, K.; Akita, M.; Inoue, K.; Takui, T.; Nakasuji, K., Four-stage rins via intramolecular oxidative coupling, Chem. Commun. 2012, amphoteric redox properties and biradicaloid character of tetra- 48, 10069-10086. tert-butyldicyclopenta[b;d]thieno[1,2,3-cd;5,6,7-c′d′]diphena- lene, Angew. Chem. Int. Ed. 2004, 43, 6474-6479.

(38) Kubo, T.; Shimizu, A.; Sakamoto, M.; Uruichi, M.; Yakushi, (57) He, Y.; Garnica, M.; Bischoff, F.; Ducke, J.; Bocquet, M.-L.; K.; Nakano, M.; Shiomi, D.; Sato, K.; Takui, T.; Morita, Y.; Nakasuji, Batzill, M.; Auwärter, W.; Barth, J. V., Fusing tetrapyrroles to gra- K., Synthesis, intermolecular interaction, and semiconductive be- phene edges by surface-assisted covalent coupling, Nat. Chem. havior of a delocalized singlet biradical hydrocarbon, Angew. 2016, 9, 33. Chem. Int. Ed. 2005, 44, 6564-6568. (58) Grill, L.; Dyer, M.; Lafferentz, L.; Persson, M.; Peters, M. V.; (39) Morita, Y.; Nishida, S. (2010). Phenalenyls, cyclopentadi- Hecht, S., Nano-architectures by covalent assembly of molecular enyls, and other carbon-centered radicals. In Stable radicals; building blocks, Nat. Nanotechnol. 2007, 2, 687. Hicks, R. G., Ed., p 81-145. (59) Auwärter, W.; Seufert, K.; Bischoff, F.; Ecija, D.; Vijayara- (40) Y. Gopalakrishna, T.; Zeng, W.; Lu, X.; Wu, J., From open- ghavan, S.; Joshi, S.; Klappenberger, F.; Samudrala, N.; Barth, J. V., shell singlet diradicaloids to polyradicaloids, Chem. Commun. A surface-anchored molecular four-level conductance switch 2018, 54, 2186-2199. based on single proton transfer, Nat. Nanotechnol. 2012, 7, 41-46. (41) Björk, J.; Hanke, F., Towards design rules for covalent (60) Rubio-Verdú, C.; Sarasola, A.; Choi, D.-J.; Majzik, Z.; nanostructures on metal surfaces, Chem. Eur. J. 2014, 20, 928-934. Ebeling, R.; Calvo, M. R.; Ugeda, M. M.; Garcia-Lekue, A.; Sánchez- (42) Lindner, R.; Kühnle, A., On-surface reactions, Chem- Portal, D.; Pascual, J. I., Orbital-selective spin excitation of a mag- PhysChem 2015, 16, 1582-1592. netic porphyrin, Commun. Phys. 2018, 1, 15. (43) Pavliček, N.; Schuler, B.; Collazos, S.; Moll, N.; Pérez, D.; (61) Cirera, B.; de la Torre, B.; Moreno, D.; Ondráček, M.; Zbořil, Guitián, E.; Meyer, G.; Peña, D.; Gross, L., On-surface generation R.; Miranda, R.; Jelínek, P.; Écija, D., On-surface synthesis of gold and imaging of arynes by , Nat. Chem. porphyrin derivatives via a cascade of chemical interactions: Pla- 2015, 7, 623. narization, self-metalation, and intermolecular coupling, Chem. (44) Kocić, N.; Liu, X.; Chen, S.; Decurtins, S.; Krejčí, O.; Jelínek, Mater. 2019, 31, 3248-3256. P.; Repp, J.; Liu, S.-X., Control of reactivity and regioselectivity for (62) Li, J.; Merino-Díez, N.; Carbonell-Sanromà, E.; Vilas- on-surface dehydrogenative aryl –aryl bond formation, J. Am. Varela, M.; de Oteyza, D. G.; Peña, D.; Corso, M.; Pascual, J. I., Sur- Chem. Soc. 2016, 138, 5585-5593. vival of spin state in magnetic porphyrins contacted by graphene (45) Held, P. A.; Fuchs, H.; Studer, A., Covalent-bond formation nanoribbons, Sci. Adv. 2018, 4, eaaq0582. via on-surface chemistry, Chem. Eur. J. 2017, 23, 5874-5892. (63) Mateo, L. M.; Sun, Q.; Liu, S.-X.; Bergkamp, J. J.; Eimre, K.; (46) Sánchez-Sánchez, C.; Nicolaï, A.; Rossel, F.; Cai, J.; Liu, J.; Pignedoli, C. A.; Ruffieux, P.; Decurtins, S.; Bottari, G.; Fasel, R.; Feng, X.; Müllen, K.; Ruffieux, P.; Fasel, R.; Meunier, V., On-surface Torres, T., On-surface synthesis and characterization of triply cyclization of ortho-dihalotetracenes to four- and six-membered fused porphyrin–graphene nanoribbon hybrids, Angew. Chem. rings, J. Am. Chem. Soc. 2017, 139, 17617-17623. Int. Ed. 2020, 59, 1334-1339. (47) Lackinger, M., Surface-assisted ullmann coupling, Chem. (64) Treier, M.; Pignedoli, C. A.; Laino, T.; Rieger, R.; Müllen, K.; Commun. 2017, 53, 7872-7885. Passerone, D.; Fasel, R., Surface-assisted cyclodehydrogenation (48) Pavliček, N.; Mistry, A.; Majzik, Z.; Moll, N.; Meyer, G.; Fox, provides a synthetic route towards easily processable and chemi- D. J.; Gross, L., Synthesis and characterization of triangulene, Nat. cally tailored nanographenes, Nat. Chem. 2010, 3, 61. Nanotechnol. 2017, 12, 308-311. (65) Lindsey, J. S.; Schreiman, I. C.; Hsu, H. C.; Kearney, P. C.; (49) Majzik, Z.; Pavliček, N.; Vilas-Varela, M.; Pérez, D.; Moll, Marguerettaz, A. M., Rothemund and adler-longo reactions revis- N.; Guitián, E.; Meyer, G.; Peña, D.; Gross, L., Studying an antiaro- ited: Synthesis of tetraphenylporphyrins under equilibrium condi- matic polycyclic hydrocarbon adsorbed on different surfaces, Nat. tions, J. Org. Chem. 1987, 52, 827-836. Commun. 2018, 9, 1198. (66) Kruszewski, J.; Krygowski, T. M., Definition of aromaticity (50) Sun, Q.; Zhang, R.; Qiu, J.; Liu, R.; Xu, W., On-surface syn- basing on the harmonic oscillator model, Tetrahedron Lett. 1972, thesis of carbon nanostructures, Adv. Mater. 2018, 30, 1705630. 13, 3839-3842. (51) Clair, S.; de Oteyza, D. G., Controlling a chemical coupling (67) Matito, E.; Poater, J.; Solá, M. (2007). Aromaticity analysis reaction on a surface: Tools and strategies for on-surface synthesis, by means of the quantum theory of atoms in molecules. In The Chem. Rev. 2019, 119, 4717-4776. quantum theory of atoms in molecules; Matta, C. F., Boyd, R. J., (52) Mishra, S.; Beyer, D.; Eimre, K.; Kezilebieke, S.; Berger, R.; Eds., p 399-423. Gröning, O.; Pignedoli, C. A.; Müllen, K.; Liljeroth, P.; Ruffieux, P.; (68) Ternes, M., Spin excitations and correlations in scanning Feng, X.; Fasel, R., Topological frustration induces unconventional tunneling spectroscopy, New Journal of Physics 2015, 17, 063016. magnetism in a nanographene, Nat. Nanotechnol. 2020, 15, 22-28. (69) Ternes, M.; Heinrich, A. J.; Schneider, W.-D., Spectroscopic (53) Li, J.; Sanz, S.; Castro-Esteban, J.; Vilas-Varela, M.; Frie- manifestations of the kondo effect on single adatoms, J. Phys.: Con- drich, N.; Frederiksen, T.; Peña, D.; Pascual, J. I., Uncovering the dens. Matter 2008, 21, 053001. triplet ground state of triangular graphene nanoflakes engineered (70) Li, J.; Sanz, S.; Corso, M.; Choi, D. J.; Peña, D.; Frederiksen, with atomic precision on a metal surface, Phys. Rev. Lett. 2020, 124, T.; Pascual, J. I., Single spin localization and manipulation in gra- 177201. phene open-shell nanostructures, Nat. Commun. 2019, 10, 200. (54) Mishra, S.; Beyer, D.; Eimre, K.; Ortiz, R.; Fernández-Ros- (71) Mugarza, A.; Krull, C.; Robles, R.; Stepanow, S.; Ceballos, sier, J.; Berger, R.; Gröning, O.; Pignedoli, C. A.; Fasel, R.; Feng, X.; G.; Gambardella, P., Spin coupling and relaxation inside mole- Ruffieux, P., Collective all-carbon magnetism in triangulene di- cule–metal contacts, Nat. Commun. 2011, 2, 490. mers, Angew. Chem. Int. Ed. 2020, 59, 12041-12047. (72) Mielke, J.; Hanke, F.; Peters, M. V.; Hecht, S.; Persson, M.; (55) Su, J.; Telychko, M.; Hu, P.; Macam, G.; Mutombo, P.; Zhang, Grill, L., Adatoms underneath single porphyrin molecules on H.; Bao, Y.; Cheng, F.; Huang, Z.-Q.; Qiu, Z.; Tan, S. J. R.; Lin, H.; au(111), J. Am. Chem. Soc. 2015, 137, 1844-1849. Jelínek, P.; Chuang, F.-C.; Wu, J.; Lu, J., Atomically precise bottom- (73) Kumar, A.; Banerjee, K.; Dvorak, M.; Schulz, F.; Harju, A.; up synthesis of π -extended [5]triangulene, Sci. Adv. 2019, 5, Rinke, P.; Liljeroth, P., Charge-transfer-driven nonplanar adsorp- eaav7717. tion of f4tcnq molecules on epitaxial graphene, ACS Nano 2017, 11, (56) Wiengarten, A.; Seufert, K.; Auwärter, W.; Auwärter, W.; 4960-4968. Ecija, D.; Diller, K.; Allegretti, F.; Bischoff, F.; Fischer, S.; Duncan, D.; Papageorgiou, A.; Klappenberger, F.; Acres, R.; Ngo, T.; Barth, J., Surface-assisted dehydrogenative homocoupling of porphine molecules, J. Am. Chem. Soc. 2014, 136, 9346-9354.